Download as pdf or txt
Download as pdf or txt
You are on page 1of 441

MAT218: Lecture Notes on Partial Differential

Equations

Steve Shkoller
Department of Mathematics
University of California
Davis, CA 95616

June 7, 2012
Contents

1 Introduction 1
1.1 Preliminaries . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Examples of PDEs . . . . . . . . . . . . . . . . . . . . . . . . 2
1.2.1 Equation of continuity . . . . . . . . . . . . . . . . . . 2
1.2.2 The heat equation . . . . . . . . . . . . . . . . . . . . 3
1.2.3 The Laplace equation . . . . . . . . . . . . . . . . . . 4
1.2.4 The 1-d wave equation . . . . . . . . . . . . . . . . . . 4
1.2.5 Equilibrium of a membrane . . . . . . . . . . . . . . . 6
1.2.6 Equation for vibrating membrane . . . . . . . . . . . . 8
1.2.7 The Euler equations . . . . . . . . . . . . . . . . . . . 8
1.2.8 Minimal surface . . . . . . . . . . . . . . . . . . . . . 10

2 Introduction to Sobolev Spaces 11


2.1 Lp spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.1.1 Notation . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.1.2 Definitions and basic properties . . . . . . . . . . . . . 12
2.1.3 Basic inequalities . . . . . . . . . . . . . . . . . . . . . 13
2.1.4 The space (Lp (X), k · kLp (X) is complete . . . . . . . 15
2.1.5 Convergence criteria for Lp functions . . . . . . . . . . 16
2.1.6 The space L∞ (X) . . . . . . . . . . . . . . . . . . . . 17
2.1.7 Approximation of Lp (X) by simple functions . . . . . 19
2.1.8 Approximation of Lp (Ω) by continuous functions . . . 19
2.1.9 Approximation of Lp (Ω) by smooth functions . . . . . 20
2.1.10 Continuous linear functionals on Lp (X) . . . . . . . . 22
2.1.11 A theorem of F. Riesz . . . . . . . . . . . . . . . . . . 23

2
CONTENTS 3

2.1.12 Weak convergence . . . . . . . . . . . . . . . . . . . . 26


2.1.13 Integral operators . . . . . . . . . . . . . . . . . . . . 29
2.1.14 Appendix 1: The monotone and dominated conver-
gence theorems and Fatou’s lemma . . . . . . . . . . . 31
2.1.15 Appendix 2: The Fubini and Tonelli Theorems . . . . 32
2.1.16 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . 33
2.2 The Sobolev spaces H k (Ω) for integers k ≥ 0 . . . . . . . . . 36
2.2.1 Weak derivatives . . . . . . . . . . . . . . . . . . . . . 36
2.2.2 Definition of Sobolev Spaces . . . . . . . . . . . . . . . 38
2.2.3 A simple version of the Sobolev embedding theorem . 40
2.2.4 Approximation of W k,p (Ω) by smooth functions . . . . 41
2.2.5 Hölder Spaces . . . . . . . . . . . . . . . . . . . . . . . 42
2.2.6 Morrey’s inequality . . . . . . . . . . . . . . . . . . . . 43
2.2.7 The Gagliardo-Nirenberg-Sobolev inequality . . . . . . 48
2.2.8 Local coordinates near ∂Ω . . . . . . . . . . . . . . . . 53
2.2.9 Sobolev extensions and traces. . . . . . . . . . . . . . 54
2.2.10 The subspace W01,p (Ω) . . . . . . . . . . . . . . . . . . 56
2.2.11 Weak solutions to Dirichlet’s problem . . . . . . . . . 58
2.2.12 Strong compactness . . . . . . . . . . . . . . . . . . . 60
2.2.13 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . 64
2.3 The Fourier Transform . . . . . . . . . . . . . . . . . . . . . . 68
2.3.1 Fourier transform on L1 (Rn ) and the space S(Rn ) . . 68
2.3.2 The topology on S(Rn ) and tempered distributions . . 72
2.3.3 Fourier transform on S 0 (Rn ) . . . . . . . . . . . . . . . 73
2.3.4 The Fourier transform on L2 (Rn ) . . . . . . . . . . . . 74
2.3.5 Bounds for the Fourier transform on Lp (Rn ) . . . . . . 75
2.3.6 Convolution and the Fourier transform . . . . . . . . . 76
2.3.7 An explicit computation with the Fourier Transform . 77
2.3.8 Applications to the Poisson, Heat, and Wave equations 79
2.3.9 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . 84
2.4 The Sobolev Spaces H s (Rn ), s ∈ R . . . . . . . . . . . . . . . 93
2.4.1 H s (Rn ) via the Fourier Transform . . . . . . . . . . . 93
4 CONTENTS

2.4.2 Fractional-order Sobolev spaces via difference quotient


norms . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
2.5 Fractional-order Sobolev spaces on domains with boundary . 103
2.5.1 The space H s (Rn+ ) . . . . . . . . . . . . . . . . . . . . 103
2.5.2 The Sobolev space H s (Ω) . . . . . . . . . . . . . . . . 104
2.6 The Sobolev Spaces H s (Tn ), s ∈ R . . . . . . . . . . . . . . . 106
2.6.1 The Fourier Series: Revisited . . . . . . . . . . . . . . 106
2.6.2 The Poisson Integral Formula and the Laplace operator109
2.6.3 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . 112
2.7 Regularity of the Laplacian on Ω . . . . . . . . . . . . . . . . 113

3 Fourier Series and its Applications 119


3.1 Inner product spaces and Hilbert spaces . . . . . . . . . . . . 119
3.2 The Hilbert space L2 (T) . . . . . . . . . . . . . . . . . . . . . 122
3.2.1 Approximations of the identity . . . . . . . . . . . . . 122
3.2.2 Trigonometric polynomials . . . . . . . . . . . . . . . 124
3.2.3 Fourier representation of functions on [0, π] . . . . . . 126
3.3 Pointwise and uniform convergence of the Fourier series . . . 127
3.3.1 Pointwise convergence . . . . . . . . . . . . . . . . . . 128
3.3.2 Uniform convergence . . . . . . . . . . . . . . . . . . . 129
3.3.3 Jump discontiunity and Gibbs phenomenon . . . . . . 135
3.4 The Sobolev space H s (T) . . . . . . . . . . . . . . . . . . . . 136
3.4.1 Characterization of H 1 (T) . . . . . . . . . . . . . . . . 138
3.4.2 The Fourier representation of f 0 for f ∈ H 1 (T) . . . . 139
3.5 1-D heat equations with periodic boundary condition . . . . . 141
3.5.1 Formal approach . . . . . . . . . . . . . . . . . . . . . 141
3.5.2 Rigorous approach . . . . . . . . . . . . . . . . . . . . 142
3.5.3 The special case f = 0 . . . . . . . . . . . . . . . . . . 147
3.5.4 Decay estimates . . . . . . . . . . . . . . . . . . . . . 148
3.6 1-D heat equations with Dirichlet boundary condition . . . . 149
3.7 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149

4 The Laplace and Poisson Equations 155


4.1 The fundamental solution . . . . . . . . . . . . . . . . . . . . 155
CONTENTS 5

4.1.1 Uniform and Hölder Continuous Functions . . . . . . 157


4.1.2 The Poisson equation −∆u = f in Rn . . . . . . . . . 159
4.2 Green’s function . . . . . . . . . . . . . . . . . . . . . . . . . 163
4.3 Properties of harmonic functions . . . . . . . . . . . . . . . . 165
4.3.1 Mean-value property . . . . . . . . . . . . . . . . . . . 165
4.3.2 Maximum principles . . . . . . . . . . . . . . . . . . . 166
4.3.3 The Harnack inequality . . . . . . . . . . . . . . . . . 167
4.3.4 Local estimates . . . . . . . . . . . . . . . . . . . . . . 168
4.3.5 Regularity of weakly harmonic functions . . . . . . . . 168
4.3.6 Liouville’s Theorem . . . . . . . . . . . . . . . . . . . 169
4.3.7 Analyticity . . . . . . . . . . . . . . . . . . . . . . . . 170
4.4 Computing Green’s functions . . . . . . . . . . . . . . . . . . 170
4.4.1 The case Ω = Rn−1 × R+ . . . . . . . . . . . . . . . . 171
4.4.2 The case Ω = B(0, 1) or Ω = B(0, R) . . . . . . . . . . 173
4.5 Perron’s method and solutions to the Poisson problem . . . . 174
4.6 Exercise . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 178

5 Second Order Linear Elliptic Equations 181


5.1 Strong solutions . . . . . . . . . . . . . . . . . . . . . . . . . . 181
5.1.1 The Dirichlet problem . . . . . . . . . . . . . . . . . . 183
5.1.2 The Neumann problem . . . . . . . . . . . . . . . . . 183
5.1.3 The Robin problem . . . . . . . . . . . . . . . . . . . 183
5.1.4 A non-standard elliptic problem . . . . . . . . . . . . 184
5.1.5 A vector-valued Dirichlet problem: Linear Elasticity . 184
5.2 Weak solutions . . . . . . . . . . . . . . . . . . . . . . . . . . 186
5.2.1 The Dirichlet problem . . . . . . . . . . . . . . . . . . 189
5.2.2 The Neumann problem . . . . . . . . . . . . . . . . . 189
5.2.3 The Robin problem . . . . . . . . . . . . . . . . . . . 189
5.2.4 A non-standard elliptic problem . . . . . . . . . . . . 190
5.2.5 Linear elasticity . . . . . . . . . . . . . . . . . . . . . 191
5.3 The Lax-Milgram theorem . . . . . . . . . . . . . . . . . . . . 191
5.4 The BBL inf-sup condition . . . . . . . . . . . . . . . . . . . 193
5.5 Existence and uniqueness of weak solutions . . . . . . . . . . 197
5.5.1 A few Preliminary Lemmas . . . . . . . . . . . . . . . 197
6 CONTENTS

5.5.2 The Dirichlet problem . . . . . . . . . . . . . . . . . . 199


5.5.3 The Neumann problem . . . . . . . . . . . . . . . . . 201
5.5.4 The Robin problem . . . . . . . . . . . . . . . . . . . 202
5.5.5 The non-standard elliptic problem . . . . . . . . . . . 203
5.5.6 The elasticity equations . . . . . . . . . . . . . . . . . 204
5.6 Weak solutions with distributional forcing functions . . . . . 206
5.7 General second order elliptic operators . . . . . . . . . . . . . 211
5.8 Horizontal convolution-by-layers and commutation estimates 215
5.8.1 Horizontal convolution-by-layers . . . . . . . . . . . . 215
5.8.2 Commutation estimates . . . . . . . . . . . . . . . . . 216
5.9 Elliptic regularity . . . . . . . . . . . . . . . . . . . . . . . . . 217
5.9.1 Dirichlet problem with homogeneous boundary condi-
tion . . . . . . . . . . . . . . . . . . . . . . . . . . . . 217
5.9.2 Dirichlet problem with inhomogeneous boundary con-
dition . . . . . . . . . . . . . . . . . . . . . . . . . . . 223
5.9.3 Neumann problem . . . . . . . . . . . . . . . . . . . . 223
5.9.4 The presence of b and c in the elliptic operator L . . . 224
5.9.5 Higher regularity . . . . . . . . . . . . . . . . . . . . . 225
5.10 Two-phase problems - Elliptic equations with mixed type of
boundary condition . . . . . . . . . . . . . . . . . . . . . . . . 226
5.10.1 The variational formulation and weak solutions . . . . 228
5.10.2 Regularity . . . . . . . . . . . . . . . . . . . . . . . . . 229
5.11 Maximum principle . . . . . . . . . . . . . . . . . . . . . . . . 229
5.12 Eigenvalues and eigenfunctions . . . . . . . . . . . . . . . . . 231
5.13 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 235

6 Linear Parabolic Equations 241


6.1 Spaces involving time . . . . . . . . . . . . . . . . . . . . . . 241
6.1.1 Compactness theorems . . . . . . . . . . . . . . . . . . 245
6.2 Second-order parabolic equations . . . . . . . . . . . . . . . . 247
6.2.1 Weak solutions . . . . . . . . . . . . . . . . . . . . . . 248
6.2.2 Existence of weak solutions - Galerkin schemes . . . . 249
6.2.3 Uniqueness of the weak solution . . . . . . . . . . . . 255
6.2.4 Regularity . . . . . . . . . . . . . . . . . . . . . . . . . 256
CONTENTS 7

6.2.5 Higher regularity and compatibility conditions . . . . 259


6.3 Parabolic equations with special forcing functions . . . . . . . 264
6.4 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 267

7 Linear Hyperbolic Equations 270


7.1 Weak solutions . . . . . . . . . . . . . . . . . . . . . . . . . . 270
7.1.1 Uniqueness of the weak solution . . . . . . . . . . . . 273
7.2 The method of parabolic regularization - the simple case . . . 275
7.2.1 The existence of the weak solution . . . . . . . . . . . 275
7.2.2 Regularity theory with smooth data . . . . . . . . . . 281
7.2.3 Regularity theory with data possessing limited regularity291
7.3 The parabolic regularization of general hyperbolic equations . 303
7.3.1 Estimates for the regularized quantities . . . . . . . . 306
7.3.2 Main theorem . . . . . . . . . . . . . . . . . . . . . . . 314
7.4 An alternative approach - the Galerkin method . . . . . . . . 315
7.4.1 Existence of weak solutions . . . . . . . . . . . . . . . 315
7.4.2 Regularity theory revisited . . . . . . . . . . . . . . . 319
7.5 Hyperbolic equations with Dirichlet boundary condition . . . 326
7.6 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 327

8 More Topics on Elliptic Equations 332


8.1 Elliptic equations with Sobolev class coefficients . . . . . . . . 332
8.2 Elliptic decomposition using the divergence and curl . . . . . 336
8.2.1 An H 1 (Ω) elliptic estimate . . . . . . . . . . . . . . . 336
8.2.2 An H 2 (Ω) elliptic estimate . . . . . . . . . . . . . . . 338
8.3 The problem of divu = f in Ω with u = 0 on ∂Ω . . . . . . . 339
8.3.1 Inequalities associated with the normal and tangential
decomposition of vector fields on ∂Ω . . . . . . . . . . 341
8.4 Vector fields with given vorticity, divergence, and normal trace348
8.4.1 The case Ω = R3+ . . . . . . . . . . . . . . . . . . . . . 348
8.4.2 The case Ω is a bounded smooth domain . . . . . . . 350
8.5 Compensated compactness and the div-curl lemma . . . . . . 353
8.6 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 356
8 CONTENTS

9 Fluid dynamics 358


9.1 The equations . . . . . . . . . . . . . . . . . . . . . . . . . . . 358
9.2 Incompressible Euler equations on a fixed domain . . . . . . . 359
9.2.1 Flow on a bounded and fixed domain . . . . . . . . . 359
9.2.2 The space of divergence-free vector fields on Ω . . . . 360
9.2.3 The Lagrangian flow of a velocity field . . . . . . . . . 362
9.2.4 Existence and uniqueness of solutions of the incom-
pressible Euler equations on fixed and bounded domain365
9.3 The Stokes equations . . . . . . . . . . . . . . . . . . . . . . . 369
9.3.1 The steady state case . . . . . . . . . . . . . . . . . . 369
9.4 Incompressible Euler with free-boundary . . . . . . . . . . . . 385
9.4.1 The Eulerian description . . . . . . . . . . . . . . . . . 385
9.4.2 The Lagrangian vorticity equation . . . . . . . . . . . 388
9.4.3 Notation . . . . . . . . . . . . . . . . . . . . . . . . . . 388
9.4.4 Properties of the cofactor matrix a, and a polynomial-
type inequality . . . . . . . . . . . . . . . . . . . . . . 389
9.4.5 Horizontal convolution-by-layers and commutation es-
timates . . . . . . . . . . . . . . . . . . . . . . . . . . 390
9.4.6 An asymptotically consistent κ-approximation of the
Euler equations . . . . . . . . . . . . . . . . . . . . . . 392
9.4.7 Construction of smooth solutions to κ-approximate
Euler equations . . . . . . . . . . . . . . . . . . . . . . 394
9.4.8 Asymptotic estimates which are independent of the
smooth parameter κ . . . . . . . . . . . . . . . . . . . 396
9.4.9 Proof of the Main Theorem . . . . . . . . . . . . . . . 408
9.4.10 Time of existence and bounds independent of κ and
existence of solutions to (9.58) . . . . . . . . . . . . . 408
9.4.11 The limit as κ → 0 . . . . . . . . . . . . . . . . . . . . 409

A A Brief Review on Analysis 411


A.1 Metric spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . 411
A.2 Linear spaces (Vector spaces) . . . . . . . . . . . . . . . . . . 412
A.3 Normed spaces . . . . . . . . . . . . . . . . . . . . . . . . . . 413
A.4 Completeness . . . . . . . . . . . . . . . . . . . . . . . . . . . 413
CONTENTS 9

A.5 Completion of a metric space . . . . . . . . . . . . . . . . . . 413


A.6 Bounded linear maps between normed linear spaces . . . . . . 414
A.7 Convolutions and integral operators . . . . . . . . . . . . . . 415
A.8 The dual space and weak topology . . . . . . . . . . . . . . . 416
A.9 Bounded linear functional on Lp (Ω) . . . . . . . . . . . . . . 417

B Important Topics in Functional Analysis 419


B.1 The Hahn-Banach theorem . . . . . . . . . . . . . . . . . . . 419
B.2 The open mapping and closed graph theorem . . . . . . . . . 421
B.3 Compact operators . . . . . . . . . . . . . . . . . . . . . . . . 421
B.3.1 Symmetric operators on Hilbert Spaces . . . . . . . . 430
Chapter 1

Introduction

1.1 Preliminaries

Theorem 1.1 (Divergence Theorem). Let Ω ⊆ Rn be a Lipschitz do-


main, that is, ∂Ω locally is the graph of a Lipschitz function, and w =
(w1 , · · · , wn ) ∈ C 1 (Ω) with outward pointing normal N . Then
Z Z
divwdx = w · NdS .
Ω ∂Ω

Now suppose that f is a scalar C 1 -function, and N = (N1 , · · · , Nn ) . By


setting w = f ei , where ei is the unit vector pointing to the positive xi -axis,
then the divergence theorem implies
Z Z
fxi dx = f Ni dS .
Ω ∂Ω

Suppose further that f is the product of two C 1 -functions h and g , then the
equality above implies
Z Z Z
ghxi dx = ghNi dS − gxi hdx .
Ω ∂Ω Ω
This is the multi-dimensional version of integration by parts.
Let Ω be a domain for which the divergence theorem holds and let u ∈
C 2 (Ω) and v ∈ C 1 (Ω)-functions. Then we have Green’s first identity:
∂u
Z Z Z Z
Dv · Dudx + v∆udx = div(vDu)dx = v dS . (1.1)
Ω Ω Ω ∂Ω ∂N

Suppose v ∈ C 2 (Ω) as well. Interchanging u and v in (1.1) and forming the


difference of the two equalities, we obtain Green’s second identity:
∂u ∂v i
Z Z h
(v∆u − u∆v)dx = v −u dS (1.2)
Ω ∂Ω ∂N ∂N

1
2 CHAPTER 1. INTRODUCTION

1.2 Examples of PDEs


1.2.1 Equation of continuity

Let u be the concentration of some physical quantity (u = u(x, t)) in a


domain Ω ⊆ Rn , and let F be the flux of the quantity, that is, F · NdS is
the rate of the quantity that passes through an area dS in the direction N
(outward pointing) normal to dS . Then

d
Z Z Z
udx = − F · NdS + qdx for all U ⊆ Ω ,
dt U ∂U U

where q is the strength of sources for the quantity. If u is smooth, by the


divergence theorem,
Z Z Z h i
ut dx = (q − divF )dx ⇒ ut + divF − q dx = 0
U U U

for all Lipschitz domain U . We then obtain the equation of continuity


ut + divF = q .

1. The conservation of mass: let %(x, t) and u(x, t) denote the density
and the velocity of a fluid at point x at time t . Then the flux F = ρu ,
and the equation of continuity reads

%t + div(ρu) = 0 ∀x ∈ Ω,t ∈ R. (1.3)

2. The traffic flow: let Ω = R (be the highway), and u denote the car den-
sity (given in vehicles per unit length). Then the flux F is a function
of u with the property that

(a) F (u) = 0 if u = 0 or u > L ,

(b) F 0 (u) > 0 if u ∈ (0, umax ) , and

(c) F 0 (u) < 0 if u ∈ (umax , L) .

If F is differential, and F 0 (u) = c(u) . Then the equation of continuity


reads

ut + c(u)ux = q ∀x ∈ R,t ∈ R.
1.2. EXAMPLES OF PDES 3

1.2.2 The heat equation

Let u(t, x) defined on (0, T ] × Ω be the temperature of a material body at


point x ∈ Ω at time t ∈ (0, T ] , and c(x) , %(x) , k(x) be the specific heat,
density, and the inner thermal conductivity of the material body at x . Then
by the conservation of heat, for any open set U ⊆ Ω ,
d
Z Z
c(x)%(x)u(t, x)dx = k(x)Du · NdS , (1.4)
dt U ∂U

where N denotes the outward-pointing unit normal of U . Assume that u


is smooth, and U is a Lipschitz domain. By the divergence theorem, (1.4)
implies
Z Z
c(x)%(x)ut (t, x)dx = div(k(x)Du)dx .
U U

Since U is arbitrary, the equation above implies

c(x)%(x)ut (t, x) − div(k(x)Du(t, x)) = 0 ∀ x ∈ Ω , t ∈ (0, T ].

If k is constant, then
X ∂2u n
c%
ut = ∆u ≡ .
k ∂x2i
i=1

If furthermore c and % are constants, then after rescaling of time we have

ut = ∆u . (1.5)

This is the standard heat equation (the prototype equation of parabolic


equations).
We need complementary conditions to specify a particular solution of
(1.5):

1. Initial condition: u(0, x) = ϕ(x) , ϕ(x) is a given function.

2. Boundary condition: if Ω is bounded, some boundary condition of u


at x ∈ ∂Ω for all time have to be introduced by physical reason to
specify a unique solution.

(a) Dirichlet condition: u(t, x) = ψ(t, x) for all x ∈ ∂Ω , t ≥ 0 , where


ψ is a given function.
4 CHAPTER 1. INTRODUCTION

∂u
(b) Neumann condition: = g for all x ∈ ∂Ω , t ≥ 0 , where g is a
∂N
given function.
∂u
(c) Robin condition: + hu = g for all x ∈ ∂Ω , t ≥ 0 , where h
∂N
and g are given functions.

1.2.3 The Laplace equation

The stationary solution (meaning that the solution is time independent) of


heat equation satisfies ∆u = 0 (the prototype of elliptic equations). Three
type of problems:

∆u = 0 in Ω ,
1. The Dirichlet problem:
u=ψ on ∂Ω .

∆u = 0 in Ω,
(
2. The Neumann problem: ∂u
=g on ∂Ω .
∂N
∆u = 0 in Ω,
(
3. The Robin problem: ∂u
+ hu = g on ∂Ω .
∂N

1.2.4 The 1-d wave equation

1. From Hooke’s law: imagine an array of little weights of mass m inter-


connected with massless springs of length h , and the springs have a
stiffness of k .

k k
m m m
u(x − h) u(x) u(x + h)

Here u(x) measures the distance from the equilibrium of the mass
situated at x . The forces exerted on the mass m at the location x are:
∂2u
FNewton = ma = m (x, t)
∂t2
FHooke = k[u(x + h, t) − u(x, t)] − k[u(x, t) − u(x − h, t)]
= k[u(x + h, t) − 2u(x, t) + u(x − h, t)] .
1.2. EXAMPLES OF PDES 5

If the array of weights consists of N weights spaced evenly over the


length L = N h of total mass M = N m , and the total stiffness of the
array K = k/N , then

∂2u KL2 u(x + h, t) − 2u(x, t) + u(x − h, t)


(x, t) = .
∂t2 M h2
Taking the limit N → ∞ , h → 0 (and assuming smoothness) we
obtain

utt (x, t) = c2 uxx (x, t) . (1.6)

(2) Equation of vibrating string: let u(x, t) measure the distance of a


string from its equilibrium.

α β
T2 String
T1 u(x) u(x + h)

x x+h

Assuming only motion in the vertical direction, the horizontal compo-


nent of tensions T1 and T2 have to be the same:

T1 cos α = T2 cos β ≈ T .

The difference of the vertical component of T1 and T2 induces the


motion in the vertical direction:
∂2u
m (x, t) ≈ T2 sin β − T1 sin α = (T2 cos β) tan β − (T1 cos α) tan α
∂t2
≈ [T (x + h)ux (x + h, t) − T (x)ux (x, t)] .

If µ is the density of the string, then m = µh ; hence

∂2u T (x + h)ux (x + h, t) − T (x)ux (x, t)


µ 2
(x, t) ≈ .
∂t h
6 CHAPTER 1. INTRODUCTION

Taking the limit h → 0 , we obtain

µutt (x, t) = [T (x)ux (x, t)]x . (1.7)

If there is an external forcing f acting on the string, then (1.7) becomes

µutt (x, t) = [T (x)ux (x, t)]x + f . (1.8)

If T is constant, then (1.8) reduces to


f
utt (x, t) = c2 uxx (x, t) + . (1.9)
µ

u(x, 0) = ϕ(x) ,
Initial conditions: , where ϕ and ψ are given functions.
ut (x, 0) = ψ(x) ,

Boundary conditions:

1. Vibration string with fixed ends: u(0, t) = u(L, t) = 0 .


∂u ∂u
2. Vibration string with free ends: (0, t) = (L, t) = 0 .
∂x ∂x

1.2.5 Equilibrium of a membrane

Let T denote the tension of a membrane, and z = u(x1 , x2 ) be the membrane


in equilibrium.

dS

dA

The work done by the inner elastic force is equal to


dS
Z  
E= T − 1 dA .
Ω dA
1.2. EXAMPLES OF PDES 7

In other words, to deform a membrane from its unforced equilibrium state


to a surface S requires the input of the energy shown above. Since the
membrane S is parametrized by (x1 , x2 , u(x1 , x2 )) for (x1 , x2 ) ∈ Ω ,
Z hp i
E= T 1 + |Du|2 − 1 dx .

Now suppose that the deformation of the membrane is very small, or more
precisely, |Du|  1 , then by the Taylor expansion,
T
Z
E≈ |Du|2 dx .
Ω 2
If the deformation of the membrane is due to a small external force f , then
the potential energy stored in the membrane is
T
Z h i
E(u) = − |Du|2 + f u dx .
Ω 2
A membrane in equilibrium state (under the action of force f ) will minimize
the potential energy.
For ϕ ∈ C 1 (Ω) , let
E(u + tϕ) − E(u)
Z h i
δE(u; ϕ) = lim = − T Du · Dϕ + f ϕ dx .
t→0 t Ω

If u ∈ C 2 (Ω) , we may integrate by parts to obtain


∂u
Z h i Z
δE(u; ϕ) = div(T Du) + f ϕdx − Tϕ dS .
Ω ∂Ω ∂N
1. Membrane fastened on the boundary: then u satisfies δE(u; ϕ) = 0 for
all ϕ ∈ Cc1 (Ω) and u = g on ∂Ω , that is,
Z h i
div(T Du) + f ϕdx = 0 ∀ ϕ ∈ Cc1 (Ω) .

Therefore, div(T Du) + f = 0 in Ω . So the problem becomes



div(T Du) + f = 0 in Ω ,
(D)
u=g on ∂Ω .

2. Membrane with free boundary: then δE(u; ϕ) = 0 for all ϕ ∈ C 1 (Ω) ,


that is,
∂u
Z h i Z
div(T Du) + f ϕdx − Tϕ dS = 0 ∀ ϕ ∈ C 1 (Ω) .
Ω ∂Ω ∂N
8 CHAPTER 1. INTRODUCTION

In particular, consider ϕ ∈ Cc1 (Ω) , then we still obtain

div(T Du) + f = 0 in Ω;
∂u ∂u
Z
hence dS = 0 for all ϕ ∈ C 1 (Ω) . Therefore,
Tϕ = 0 on
∂Ω ∂N ∂N
∂Ω , and the problem becomes

 div(T Du) + f = 0 in Ω ,
(N) ∂u
 =0 on ∂Ω .
∂N

1.2.6 Equation for vibrating membrane

Let T be the tension, % be the density, and f be the density of the external
force which may depend on x and t .
d’Alembert’s principle:
Z h i
− T Du · Dϕ + (f − %utt )ϕ dx = 0

for all ϕ compatible with the existence constraints. Therefore,

1. Membrane fastened on the boundary:




 %utt − div(T Du) = f in Ω × (0, T ] ,

u=g on ∂Ω × (0, T ] ,


u(x, 0) = g(x), ut (x, 0) = h(x) for all x ∈ Ω .

2. Membrane with free boundary:




 %utt − div(T Du) = f in Ω × (0, T ] ,

∂u

=0 on ∂Ω × (0, T ] ,


 ∂N
u(x, 0) = g(x), ut (x, 0) = h(x) for all x ∈ Ω .

1.2.7 The Euler equations

Aside from the equation of continuity (1.3), at least an equation for the
velocity u is required to complete the system. Consider that conservation of
momentum. Let m = %u be the momentum. The conservation of momentum
states that
d
Z Z Z Z
mdx = − m(u · N)dS − pNdS + %f dx ,
dt U ∂U ∂U U
1.2. EXAMPLES OF PDES 9

here we use the fact that the rate of change of momentum of a body is equal
to the resultant force acting onZthe body, and with p denoting the pressure
the buoyancy force is given by pNdS . Here we assume that the fluid is
∂U
invicid so that no friction force is presented in the fluid. Therefore, assuming
the smoothness of the variables, the divergence theorem implies that
n
∂(muj )
Z h X i
mt + + Dp − %f dx = 0 for all Lipschitz domain U ⊆ Ω .
U ∂xj
j=1

As a consequence, we obtain the momentum equation

(%u)t + div(%u ⊗ u) = −Dp + %f ∀x ∈ Ω,t ∈ R. (1.10)

Initial conditions: %(x, 0) = %0 (x) and u(x, 0) = u0 (x) for all x ∈ Ω .


Boundary condition: u · N = 0 on ∂Ω .

1. If the density is constant (such as water), then (1.3) and (1.10) reduce
to

ut + u · Du = −Dp + f in Ω × (0, T ) , (1.11a)


divu = 0 in Ω × (0, T ) . (1.11b)

Equation (1.11) together with the initial and the boundary condition
are called the incompressible Euler equations.

2. If the pressure p solely depends on the density, that is, p = p(%) (the
equation of state), then (1.3) and (1.10) together with are called the
isentropic Euler equations.

3. If the pressure p is also a function of the temperature (see Boyle’s law


for example), that is, p = p(%, ϑ) where ϑ denotes the temperature,
then an equation for ϑ is also required to complete the system.

4. Consider a very slow motion of the fluid, that is, |u| + |Du|  1 . Sup-
pose p0 and %0 are the pressure and density of the fluid in equilibrium.
Let % = %0 + %̃ , p = p0 + p̃ , where %̃ and p̃ are small. Then
1
(a) ut ≈ − Dp̃ , (b) %̃t + %0 divu ≈ 0 , (c) p̃ ≈ p 0 (%0 )%̃ .
%0
10 CHAPTER 1. INTRODUCTION

Using (c) in (b), we find that


1
(b)’ p̃t ≈ −p 0 (%0 )divu .
%0
Suppose now the fluid motion is rotation free, then u = Dϕ for some
scalar function ϕ called the velocity potential of the fluid motion. Then
(a) implies p̃ = −%0 ϕt , hence (b)’ implies

ϕtt = c2 div(Dϕ) = c2 ∆ϕ ,
p
where c2 = p 0 (%0 ) is the sound speed.

1.2.8 Minimal surface

Suppose that Ω ⊆ R2 is a bounded set with boundary parametrized by


(x(t), y(t)) for t ∈ I . Let C ⊆ R3 be a closed curve parametrized by
(x(t), y(t), f (x(t), y(t))) . We want to find a surface having C as its boundary
with minimal surface area. Then the goal is to find a function u with the
property that u = f on ∂Ω that minimizes the functional
Z p
A(w) = 1 + |Dw|2 dA .

Let ϕ ∈ C 1 (Ω) , and define

A(u + tϕ) − A(u) Du · Dϕ


Z
δA(u; ϕ) = lim = p dx .
t→0 t Ω 1 + |Du|2

If u minimize A , then δA(u; ϕ) = 0 for all ϕ ∈ Cc1 (Ω) . Assuming that


u ∈ C 2 (Ω) , we find that u satisfies
 Du 
div p = 0,
1 + |Du|2

or expanding the bracket using the Lebnitz rule, we obtain the minimal
surface equation

(1 + u2y )uxx − 2ux uy uxy + (1 + u2x )uyy = 0 ∀ (x, y) ∈ Ω . (1.12)


Chapter 2

Introduction to Sobolev
Spaces

2.1 Lp spaces

2.1.1 Notation

We will usually use Ω to denote an open and smooth domain in Rd , for


d = 1, 2, 3, ... In this chapter on Lp spaces, we will sometimes use X to
denote a more general measure space, but the reader can usually think of a
subset of Euclidean space.

C k (Ω) is the space of functions which are k times differentiable in Ω for


integers k ≥ 0.

C 0 (Ω) then coincides with C(Ω), the space of continuous functions on Ω.

C ∞ (Ω) = ∩k≥0 C k (Ω).

spt f denotes the support of a function f , and is the closure of the set
{x ∈ Ω | f (x) 6= 0}.

C0 (Ω) = {u ∈ C(Ω) | spt u compact in Ω}.

C0k (Ω) = C k (Ω) ∩ C0 (Ω).


C0∞ (Ω) = C ∞ (Ω)∩C0 (Ω). We will also use D(Ω) to denote this space, which
is known as the space of test functions in the theory of distributions.

11
12 CHAPTER 2. INTRODUCTION TO SOBOLEV SPACES

2.1.2 Definitions and basic properties

Definition 2.1. Let 0 < p < ∞ and let (X, M, µ) denote a measure space.
If f : X → R is a measurable function, then we define

Z 1
p
p
kf kLp (X) := |f | dx and kf kL∞ (X) := ess supx∈X |f (x)| .
X

Note that kf kLp (X) may take the value ∞. Unless stated otherwise, we will
usually consider X to be a smooth, open subset Ω of Rd , and we will assume
that all functions under consideration are measurable.

Definition 2.2. The space Lp (X) is the set

Lp (X) = {f : X → R | kf kLp (X) < ∞} .

The space Lp (X) satisfies the following vector space properties:

1. For each α ∈ R, if f ∈ Lp (X) then αf ∈ Lp (X);

2. If f, g ∈ Lp (X), then

|f + g|p ≤ 2p−1 (|f |p + |g|p ) ,

so that f + g ∈ Lp (X).

3. The triangle inequality is valid if p ≥ 1.

The most interesting cases are p = 1, 2, ∞, while all of the Lp arise often in
nonlinear estimates.

Definition 2.3. The space lp , called “little Lp ”, will be useful when we


introduce Sobolev spaces on the torus and the Fourier series. For 1 ≤ p < ∞,
we set

( )
X
lp = {xn }n∈Z | |xn |p < ∞ ,
n=−∞

where Z denotes the integers.


2.1. LP SPACES 13

2.1.3 Basic inequalities

Convexity is fundamental to Lp spaces for p ∈ [1, ∞).

Lemma 2.4. For λ ∈ (0, 1), xλ ≤ (1 − λ) + λx.

Proof. Set f (x) = (1 − λ) + λx − xλ ; hence, f 0 (x) = λ − λxλ−1 = 0 if and


only if λ(1 − xλ−1 ) = 0 so that x = 1 is the critical point of f . In particular,
the minimum occurs at x = 1 with value

f (1) = 0 ≤ (1 − λ) + λx − xλ .

Lemma 2.5. For a, b ≥ 0 and λ ∈ (0, 1), aλ b1−λ ≤ λa + (1 − λ)b with


equality if a = b.

Proof. If either a = 0 or b = 0, then this is trivially true, so assume that


a, b > 0. Set x = a/b, and apply Lemma 1 to obtain the desired inequality.


Theorem 2.6 (Hölder’s inequality). Suppose that 1 ≤ p ≤ ∞ and 1 < q <


1 1
∞ with p + q = 1. If f ∈ Lp and g ∈ Lq , then f g ∈ L1 . Moreover,

kf gkL1 ≤ kf kLp kgkLq .

Note that if p = q = 2, then this is the Cauchy-Schwarz inequality since


kf gkL1 = |(f, g)L2 |.

Proof. We use Lemma 2.5. Let λ = 1/p and set


|f |p |g|q
a= , and b =
kf kpLp kgkqLp

for all x ∈ X. Then aλ b1−λ = a1/p b1−1/p = a1/p b1/q so that


|f | · |g| 1 |f |p 1 |g|q
≤ p + .
kf kLp kgkLq p kf kLp q kgkqLq
Integrating this inequality yields
|f | · |g| 1 |f |p 1 |g|q
Z  
1 1
Z
dx ≤ p + q dx = + = 1 .
X kf kL kgkL
p q
X p kf kLp q kgkLq p q

14 CHAPTER 2. INTRODUCTION TO SOBOLEV SPACES

p 1
Definition 2.7. The exponent q = p−1 (or q = 1− p1 ) is called the conjugate
exponent of p.

Theorem 2.8 (Minkowski’s inequality). If 1 ≤ p ≤ ∞ and f, g ∈ Lp then

kf + gkLp ≤ kf kLp + kgkLp .

Proof. If f + g = 0 a.e., then the statement is trivial. Assume that f + g 6= 0


a.e. Consider the equality

|f + g|p = |f + g| · |f + g|p−1 ≤ (|f | + |g|)|f + g|p−1 ,

and integrate over X to find that


Z Z
|f + g|p dx ≤ (|f | + |g|)|f + g|p−1 dx
 
X X
Hölder’s
(kf kLp + kgkLp ) |f + g|p−1 Lq .

p
Since q = p−1 ,

Z 1
q
|f + g|p−1 q = p

L
|f + g| dx ,
X

from which it follows that


Z 1− 1
q
p
|f + g| dx ≤ kf kLp + kgkLq ,
X
1
which completes the proof, since p = 1 − 1q . 

Corollary 2.9. For 1 ≤ p ≤ ∞, Lp (X) is a normed linear space.

Example 2.10. Let Ω denote a subset of Rn whose Lebesgue measure is


equal to one. If f ∈ L1 (Ω) satisfies f (x) ≥ M > 0 for almost all x ∈ Ω,
then log(f ) ∈ L1 (Ω) and satisfies
Z Z
log f dx ≤ log( f dx) .
Ω Ω

To see this, consider the function g(t) = t − 1 − log t for t > 0. Compute
g 0 (t) = 1− 1t = 0 so t = 1 is a minimum (since g 00 (1) > 0). Thus, log t ≤ t−1
1
and letting t 7→ t we see that
1
1− ≤ log t ≤ t − 1 . (2.1)
t
2.1. LP SPACES 15

Since log x is continuous and f is measurable, then log f is measurable for


f (x)
f > 0. Let t = kf kL1 in (2.1) to find that

kf kL1 f (x)
1− ≤ log f (x) − log kf kL1 ≤ − 1. (2.2)
f (x) kf kL1
Since g(x) ≤ log f (x) ≤ h(x) for two integrable functions g and h, it fol-
lows that log f (x) is integrable. Next, integrate (2.2) to finish the proof, as
R  f (x) 
X kf k 1 − 1 dx = 0.
L

2.1.4 The space (Lp (X), k · kLp (X) is complete

Recall the a normed linear space is a Banach space if every Cauchy sequence
has a limit in that space; furthermore, recall that a sequence xn → x in X
if limn→∞ kxn − xkX = 0.
The proof of completeness makes use of the following two lemmas which
are restatements of the Monotone Convergence Theorem and the Dominated
Convergence Theorem, respectively (see the Appendix for this chapter).

Lemma 2.11 (MCT). If fn ∈ L1 (X), 0 ≤ f1 (x) ≤ f2 (x) ≤ · · ·, and


kfn kL1 (X) ≤ C < ∞, then limn→∞ fn (x) = f (x) with f ∈ L1 (X) and
kfn − f kL1 → 0 as n → 0.

Lemma 2.12 (DCT). If fn ∈ L1 (X), limn→∞ fn (x) = f (x) a.e., and if


∃ g ∈ L1 (X) such that |fn (x)| ≤ |g(x)| a.e. for all n, then f ∈ L1 (X) and
kfn − f kL1 → 0.

Proof. Apply the Dominated Convergene Theorem to the sequence hn =


|fn − f | → 0 a.e., and note that |hn | ≤ 2g. 

Theorem 2.13. If 1≤ p < ∞ then Lp (X) is a Banach space.

Proof. Step 1. The Cauchy sequence. Let {fn }∞


n=1 denote a Cauchy
sequence in Lp , and assume without loss of generality (by extracting a sub-
sequence if necessary) that kfn+1 − fn kLp ≤ 2−n .

Step 2. Conversion to a convergent monotone sequence. Define the


sequence {gn }∞
n=1 as

g1 = 0, gn = |f1 | + |f2 − f1 | + · · · + |fn − fn−1 | for n ≥ 2 .


16 CHAPTER 2. INTRODUCTION TO SOBOLEV SPACES

It follows that
0 ≤ g1 ≤ g2 ≤ · · · ≤ gn ≤ · · ·

so that gn is a monotonically increasing sequence. Furthermore, {gn } is


uniformly bounded in Lp as
Z ∞
!p
gnp dx = kgn kpLp ≤ kf1 kLp +
X
kfi − fi−1 kLp ≤ (kf1 kLp + 1)p ;
X i=2

thus, by the Monotone Convergence Theorem, gnp % g p a.e., g ∈ Lp , and


gn ≤ g a.e.

Step 3. Pointwise convergence of {fn }. For all k ≥ 1,

|fn+k − fn | = |fn+k − fn+k−1 + fn+k−1 + · · · − fn+1 + fn+1 − fn |


n+k+1
X
≤ |fi − fi−1 | = gn+k − gn −→ 0 a.e.
i=n+1

Therefore, fn → f a.e. Since


n
X
|fn | ≤ |f1 | + |fi − fi−1 | ≤ gn ≤ g for all n ∈ N ,
i=2

it follows that |f | ≤ g a.e. Hence, |fn |p ≤ g p , |f |p ≤ g p , and |f − fn |p ≤ 2g p ,


and by the Dominated Convergence Theorem,
Z Z
p
lim |f − fn | dx = lim |f − fn |p dx = 0 .
n→∞ X X n→∞


2.1.5 Convergence criteria for Lp functions

If {fn } is a sequence in Lp (X) which converges to f in Lp (X), then there


exists a subsequence {fnk } such that fnk (x) → f (x) for almost every x ∈ X
(denoted by a.e.), but it is in general not true that the entire sequence itself
will converge pointwise a.e. to the limit f , without some further conditions
holding.

Example 2.14. Let X = [0, 1], and consider the subintervals


                   
1 1 1 1 2 2 1 1 2 2 3 3 1
0, , , 1 , 0, , , , , 1 , 0, , , , , , , 1 , 0, , ···
2 2 3 3 3 3 4 4 4 4 4 4 5
Let fn denote the indicator function of the nth interval of the above sequence.
Then kfn kLp → 0, but fn (x) does not converge for any x ∈ [0, 1].
2.1. LP SPACES 17

Example 2.15. Set X = R, and for n ∈ N, set fn = 1[n,n+1] . Then


fn (x) → 0 as n → ∞, but kfn kLp = 1 for p ∈ [1, ∞); thus, fn → 0
pointwise, but not in Lp .

Example 2.16. Set X = [0, 1], and for n ∈ N, set fn = n1[0, 1 ] . Then
n
fn (x) → 0 a.e. as n → ∞, but kfn kL1 = 1; thus, fn → 0 pointwise, but not
in L1 .

Theorem 2.17. For 1 ≤ p < ∞, suppose that {fn } ⊂ Lp (X) and that
fn (x) → f (x) a.e. If limn→∞ kfn kLp (X) = kf kLp (X) , then fn → f in Lp (X).

a+b p
≤ 12 (ap + bp ) so that

Proof. Given a, b ≥ 0, convexity implies that 2
(a + b)p ≤ 2p−1 (ap + bp ), and hence |a − b|p ≤ 2p−1 (|a|p + |b|p ). Set a = fn
and b = f to obtain the inequality

0 ≤ 2p−1 (|fn |p + |f |p ) − |fn − f |p .

Since fn (x) → f (x) a.e.,


Z Z
p p
lim 2p−1 (|fn |p + |f |p ) − |fn − f |p dx .

2 |f | dx =
X X n→∞

Thus, Fatou’s lemma asserts that


Z Z
p p
2p−1 (|fn |p + |f |p ) − |fn − f |p dx

2 |f | dx ≤ lim inf
X n→∞
Z X Z  Z 
p−1 p p−1 p p
=2 |f | dx + 2 lim |fn | + lim inf − |fn − f | dx
X n→∞ X n→∞ X
Z Z
p−1 p
=2 |f | dx − lim sup |fn − f |p dx .
X n→∞ X

|f |p dx < ∞, the last inequality shows that lim supn→∞ X |fn −


R R
As X
f |p dx ≤ 0. It follows that lim supn→∞ X |fn − f |p dx = lim inf n→∞ X |fn −
R R

f |p dx = 0, so that limn→∞ X |fn − f |p dx = 0.


R


2.1.6 The space L∞ (X)

Definition 2.18. With kf kL∞ (X) = inf{M ≥ 0 | |f (x)| ≤ M a.e.}, we set

L∞ (X) = {f : X → R | kf kL∞ (X) < ∞} .

Theorem 2.19. (L∞ (X), k · kL∞ (X) ) is a Banach space.


18 CHAPTER 2. INTRODUCTION TO SOBOLEV SPACES

Proof. Let fn be a Cauchy sequence in L∞ (X). It follows that |fn − fm | ≤


kfn − fm kL∞ (X) a.e. and hence fn (x) → f (x) a.e., where f is measurable
and essentially bounded.
Choose  > 0 and N () such that kfn −fm kL∞ (X) <  for all n, m ≥ N ().
Since |f (x) − fn (x)| = limm→∞ |fm (x) − fn (x)| ≤  holds a.e. x ∈ X, it
follows that kf − fn kL∞ (X) ≤  for n ≥ N (), so that kfn − f kL∞ (X) → 0. 

Remark 2.20. In general, there is no relation of the type Lp ⊂ Lq . For


1
example, suppose that X = (0, 1) and set f (x) = x− 2 . Then f ∈ L1 (0, 1),
but f 6∈ L2 (0, 1). On the other hand, if X = (1, ∞) and f (x) = x−1 , then
f ∈ L2 (1, ∞), but f 6∈ L1 (1, ∞).

Lemma 2.21 (Lp comparisons). If 1 ≤ p < q < r ≤ ∞, then (a) Lp ∩ Lr ⊂


Lq , and (b) Lq ⊂ Lp + Lr .

Proof. We begin with (b). Suppose that f ∈ Lq , define the set E = {x ∈


X : |f (x)| ≥ 1}, and write f as

f = f 1E + f 1E c
= g + h.

Our goal is to show that g ∈ Lp and h ∈ Lr . Since |g|p = |f |p 1E ≤ |f |q 1E


and |h|r = |f |r 1E c ≤ |f |q 1E c , assertion (b) is proven.
For (a), let λ ∈ [0, 1] and for f ∈ Lq ,
Z 1 Z 1
q q
q λq (1−λ)q
kf kLq = |f | dx = |f | |f | dµ
X X
 1
(1−λ)q (1−λ)
kf kλq
q
≤ Lp kf kLr = kf kλLp kf kLr .

Theorem 2.22. If µ(X) ≤ ∞ and q > p, then Lq ⊂ Lp .

Proof. Consider the case that q = 2 and p = 1. Then by the Cauchy-Schwarz


inequality, Z Z p
|f |dx = |f | · 1 dx ≤ kf kL2 (X) µ(X) .
X X

2.1. LP SPACES 19

2.1.7 Approximation of Lp (X) by simple functions


Pn
Lemma 2.23. If p ∈ [1, ∞), then the set of simple functions f = i=1 ai 1Ei ,
where each Ei is an element of the σ-algebra A and µ(Ei ) < ∞, is dense in
Lp (X, A, µ).

Proof. If f ∈ Lp , then f is measurable; thus, there exists a sequence {φn }∞


n=1
of simple functions, such that φn → f a.e. with

0 ≤ |φ1 | ≤ |φ2 | ≤ · · · ≤ |f |,

i.e., φn approximates f from below.


Recall that |φn − f |p → 0 a.e. and |φn − f |p ≤ 2p |f |p ∈ L1 , so by the
Dominated Convergence Theorem, kφn − f kLp → 0.
Now, suppose that the set Ei are disjoint; then by the definition of the
Lebesgue integral,
Z n
X
φpn dx = |ai |p µ(Ei ) < ∞ .
X i=1

If ai 6= 0, then µ(Ei ) < ∞. 

2.1.8 Approximation of Lp (Ω) by continuous functions

Lemma 2.24. Suppose that Ω ⊂ Rn is bounded. Then C 0 (Ω) is dense in


Lp (Ω) for p ∈ [1, ∞).

Proof. Let K be any compact subset of Ω. The functions

1
FK ,n (x) = ∈ C 0 (Ω) satisfy FK ,n ≤ 1 ,
1 + n dist(x, K)
and decrease monotonically to the characteristic function 1K . The Mono-
tone Convergence Theorem gives

fK ,n → 1K in Lp (Ω), 1 ≤ p < ∞ .

Next, let A ⊂ Ω be any measurable set, and let λ denote the Lebesgue
measure. Then

λ(A) = sup{µ(K) : K ⊂ A, K compact} .


20 CHAPTER 2. INTRODUCTION TO SOBOLEV SPACES

It follows that there exists an increasing sequence of Kj of compact subsets


of A such that λ(A\ ∪j Kj ) = 0. By the Monotone Convergence Theorem,
1Kj → 1A in Lp (Ω) for p ∈ [1, ∞). According to Lemma 2.23, each function
in Lp (Ω) is a norm limit of simple functions, so the lemma is proved. 

2.1.9 Approximation of Lp (Ω) by smooth functions

For Ω ⊂ Rn open, for  > 0 taken sufficiently small, define the open subset
of Ω by
Ω := {x ∈ Ω | dist(x, ∂Ω) > } .

Definition 2.25 (Mollifiers). Define η ∈ C ∞ (Rn ) by


2 −1
Ce(|x| −1)

if |x| < 1
η(x) := ,
0 if |x| ≥ 1
R
with constant C > 0 chosen such that Rn η(x)dx = 1.
For  > 0, the standard sequence of mollifiers on Rn is defined by

η (x) = −n η(x/) ,


R
and satisfy Rn η (x)dx = 1 and spt(η ) ⊂ B(0, ).

Definition 2.26. For Ω ⊂ Rn open, set

Lploc (Ω) = {u : Ω → R | u ∈ Lp (Ω̃) ∀ Ω̃ ⊂⊂ Ω} ,

where Ω̃ ⊂⊂ Ω means that there exists K compact such that Ω̃ ⊂ K ⊂ Ω.


We say that Ω̃ is compactly contained in Ω.

Definition 2.27 (Mollification of L1 ). If f ∈ L1loc (Ω), define its mollifica-


tion
f  = η ∗ f in Ω ,

so that
Z Z
f  (x) = η (x − y)f (y)dy = η (y)f (x − y)dy ∀x ∈ Ω .
Ω B(0,)

Theorem 2.28 (Mollification of Lp (Ω)).

(A) f  ∈ C ∞ (Ω ).
2.1. LP SPACES 21

(B) f  → f a.e. as  → 0.

(C) If f ∈ C 0 (Ω), then f  → f uniformly on compact subsets of Ω.

(D) If p ∈ [1, ∞) and f ∈ Lploc (Ω), then f  → f in Lploc (Ω).

Proof. Part (A). We rely on the difference quotient approximation of the


partial derivative. Fix x ∈ Ω , and choose h sufficiently small so that x +
hei ∈ Ω for i = 1, ..., n, and compute the difference quotient of f  :

f  (x + hei ) − f (x) x + hei − y x−y


    
1
Z
= −n η −η f (y)dy
h Ω h  
x + hei − y x−y
    
1
Z
−n
= η −η f (y)dy
Ω̃ h  

for some open set Ω̃ ⊂⊂ Ω. On Ω̃,

x + hei − y x−y 1 ∂η x − y
      
1
lim η −η = ,
h→0 h    ∂xi 

so by the Dominated Convergence Theorem,

∂f ∂η
Z
(x) = (x − y)f (y)dy .
∂xi Ω ∂xi

A similar argument for higher-order partial derivatives proves (A).

Step 2. Part (B). By the Lebesgue differentiation theorem,

1
Z
lim |f (y) − f (x)|dy = 0 for a.e. x ∈ Ω .
→0 |B(x, )| B(x,)

Choose x ∈ Ω for which this limit holds. Then


Z
|f (x) − f (x)| ≤ η (x − y)|f (y) − f (x)|dy
B(x,)
1
Z
= n η((x − y)/)|f (y) − f (x)|dy
 B(x,)
C
Z
≤ |f (x) − f (y)|dy −→ 0 as  → 0.
|B(x, )| B(x,)

Step 3. Part (C). For Ω̃ ⊂ Ω, the above inequality shows that if f ∈ C 0 (Ω)
and hence uniformly continuous on Ω̃, then f  (x) → f (x) uniformly on Ω̃.
22 CHAPTER 2. INTRODUCTION TO SOBOLEV SPACES

Step 4. Part (D). For f ∈ Lploc (Ω), p ∈ [1, ∞), choose open sets U ⊂⊂
D ⊂⊂ Ω; then, for  > 0 small enough,

kf  kLp (U ) ≤ kf kLp (D) .

To see this, note that


Z
|f  (x)| ≤ η (x − y)|f (y)|dy
B(x,)
Z
= η (x − y)(p−1)/p η (x − y)1/p |f (y)|dy
B(x,)
Z !(p−1)/p Z !1/p
p
≤ η (x − y)dy η (x − y)|f (y)| dy ,
B(x,) B(x,)

so that for  > 0 sufficiently small


Z Z Z
 p
|f (x)| dx ≤ η (x − y)|f (y)|p dydx
U U B(x,)
Z Z ! Z
p
≤ |f (y)| η (x − y)dx dy ≤ |f (y)|p dy .
D B(y,) D

Since C 0 (D) is dense in Lp (D), choose g ∈ C 0 (D) such that kf −


gkLp (D) < δ; thus

kf  − f kLp (U ) ≤ kf  − g  kLp (U ) + kg  − gkLp (U ) + kg − f kLp (U )


≤ 2kf − gkLp (D) + kg  − gkLp (U ) ≤ 2δ + kg  − gkLp (U ) .

2.1.10 Continuous linear functionals on Lp (X)

Let Lp (X)0 denote the dual space of Lp (X). For φ ∈ Lp (X)0 , the operator
norm of φ is defined by kφkop = supLp (X)=1 |φ(f )|.

p
Theorem 2.29. Let p ∈ (1, ∞], q = p−1 . For g ∈ Lq (X), define Fg :
Lp (X) → R as Z
Fg (f ) = f gdx .
X
Then Fg is a continuous linear functional on Lp (X) with operator norm
kFg kop = kgkLq (X) .
2.1. LP SPACES 23

Proof. The linearity of Fg again follows from the linearity of the Lebesgue
integral. Since
Z Z

|Fg (f )| = f gdx ≤ |f g| dx ≤ kf kLp kgkLq ,
X X

with the last inequality following from Hölder’s inequality, we have that
supkf kLp =1 |Fg (f )| ≤ kgkLq .
For the reverse inequality let f = |g|q−1 sgn g. f is measurable and in Lp
q
since |f |p = |f | q−1 = |g|q and since f g = |g|q ,
Z Z Z 1+1
p q
q q
Fg (f ) = f gdx = |g| dx = |g| dx
X X X
Z  1 Z 1
p q
= |f |p dx |g q | dx = kf kLp kgkLq
X X

Fg (f )
so that kgkLq = ≤ kFg kop .
kf kLp


Remark 2.30. Theorem 2.29 shows that for 1 < p ≤ ∞, there exists a
linear isometry g 7→ Fg from Lq (X) into Lp (X)0 , the dual space of Lp (X).
When p = ∞, g 7→ Fg : L1 (X) → L∞ (X)0 is rarely onto (L∞ (X)0 is strictly
larger than L1 (X)); on the other hand, if the measure space X is σ-finite,
then L∞ (X) = L1 (X)0 .

2.1.11 A theorem of F. Riesz

Theorem 2.31 (Representation theorem). Suppose that 1 < p < ∞ and


p
φ ∈ Lp (X)0 . Then there exists g ∈ Lq (X), q = p−1 such that
Z
φ (f ) = f gdx ∀f ∈ Lp (X) ,
X

and kφkop = kgkLq .

Corollary 2.32. For p ∈ (1, ∞) the space Lp (X, µ) is reflexive, i.e., Lp (X)00 =
Lp (X).

The proof Theorem 2.31 crucially relies on the Radon-Nikodym theorem,


whose statement requires the following definition.
24 CHAPTER 2. INTRODUCTION TO SOBOLEV SPACES

Definition 2.33. If µ and ν are measure on (X, A) then ν  µ if ν(E) = 0


for every set E for which µ(E) = 0. In this case, we say that ν is absolutely
continuous with respect to µ.

Theorem 2.34 (Radon-Nikodym). If µ and ν are two finite measures on


X, i.e., µ(X) < ∞, ν(X) < ∞, and ν  µ, then
Z Z
F (x) dν(x) = F (x)h(x)dµ(x) (2.3)
X X

holds for some nonnegative function h ∈ L1 (X, µ) and every positive mea-
surable function F .

Proof. Define measures α = µ+2ν and ω = 2µ+ν, and let H = L2 (X,


Z α) (a
Hilbert space) and suppose φ : L2 (X, α) → R is defined by φ (f ) = f dω.
X
We show that φ is a bounded linear functional since
Z Z Z

|φ(f )| = f d(2µ + ν) ≤
|f | d(2µ + 4ν) = 2 |f | dα
X X X
p
≤ kf kL2 (x,α) α(X) .

Thus, by the Riesz representation theorem, there exists g ∈ L2 (X, α)


such that Z Z
φ(f ) = f dω = f g dα ,
X X
which implies that
Z Z
f (2g − 1)dν = f (2 − g)dµ . (2.4)
X X
F
Given 0 ≤ F a measurable function on X, if we set f = 2g−1 and
2−g R R
h = 2g−1 then X F dν = X F h dx which is the desired result, if we can
prove that 1/2 ≤ g(x) ≤ 2. Define the sets
   
1 1 1 2 1
En = x ∈ X | g(x) < − and En = x ∈ X | g(x) > 2 + .
2 n n
By substituting f = 1Enj , j = 1, 2 in (2.4), we see that

µ(Enj ) = ν(Enj ) = 0 for j = 1, 2 ,

from which the bounds 1/2 ≤ g(x) ≤ 2 hold. Also µ({x ∈ X | g(x) =
1/2}) = 0 and ν({x ∈ X | g(x) = 2}) = 0. Notice that if F = 1, then
h ∈ L1 (X). 
2.1. LP SPACES 25

Remark 2.35. The more general version of the Radon-Nikodym theorem.


Suppose that µ(X) < ∞, ν is a finite signed measure (by the Hahn decom-
position, ν = ν − + ν + ) such that ν  µ; then, there exists h ∈ L1 (X, µ)
R R
such that X F dν = X F h dµ.

Lemma 2.36 (Converse to Hölder’s inequality). Let µ(X) < ∞. Suppose


that g is measurable and f g ∈ L1 (X) for all simple functions f . If
 Z 

M (g) = sup f g dµ : f is a simple function < ∞ , (2.5)

kf kLp =1 X

then g ∈ Lq (X), and kgkLq (X) = M (g).

Proof. Let φn be a sequence of simple functions such that φn → g a.e. and


|φn | ≤ |g|. Set
|φn |q−1 sgn (φn )
fn =
kφn kq−1
Lq
so that kfn kLp = 1 for p = q/(q − 1). By Fatou’s lemma,
Z
kgkLq (X) ≤ lim inf kφn kLq (X) = lim inf |fn φn |dµ .
n→∞ n→∞ X
Since φn → g a.e., then
Z Z
kgkLq (X) ≤ lim inf |fn φn |dµ ≤ lim inf |fn g|dµ ≤ M (g) .
n→∞ X n→∞ X
The reverse inequality is implied by Hölder’s inequality. 

Proof of the Lp (X)0 representation theorem. We have already proven that


there exists a natural inclusion ι : Lq (X) → Lp (X)0 which is an isometry. It
remains to show that ι is surjective.
Let φ ∈ Lp (X)0 and define a set function ν on measurable subsets E ⊂ X
by Z
ν(E) = 1E dν =: φ(1E ) .
X
Thus, if µ(E) = 0, then ν(E) = 0. Then
Z
f dν =: φ(f )
X

for all simple functions f , and by Lemma 2.23, this holds for all f ∈ Lp (X).
By the Radon-Nikodym theorem, there exists 0 ≤ g ∈ L1 (X) such that
Z Z
f dν = f g dµ ∀ f ∈ Lp (X) .
X X
26 CHAPTER 2. INTRODUCTION TO SOBOLEV SPACES

But Z Z
φ(f ) = f dν = f g dµ (2.6)
X X
and since φ ∈ Lp (X)0 , then M (g) given by (2.5) is finite, and by the converse
to Hölder’s inequality, g ∈ Lq (X), and kφkop = M (g) = kgkLq (X) . 

2.1.12 Weak convergence

The importance of the Representation Theorem 2.31 is in the use of the


weak-* topology on the dual space Lp (X)0 . Recall that for a Banach space

B and for any sequence φj in the dual space B0 , φj * φ in B0 weak-*, if
hφj , f i → hφ, f i for each f ∈ B, where h·, ·i denotes the duality pairing
between B0 and B.

Theorem 2.37 (Alaoglu’s Lemma). If B is a Banach space, then the closed


unit ball in B0 is compact in the weak -* topology.

Definition 2.38. For 1 ≤ p < ∞, a sequence {fn } ⊂ Lp (X) is said to


weakly converge to f ∈ Lp (X) if
p
Z Z
fn (x)φ(x)dx → f (x)φ(x)dx ∀φ ∈ Lq (X), q = .
X X p−1
We denote this convergence by saying that fn * f in Lp (X) weakly.

Given that Lp (X) is reflexive for p ∈ (1, ∞), a simple corollary of


Alaoglu’s Lemma is the following

Theorem 2.39 (Weak compactness for Lp , 1 < p < ∞). If 1 < p < ∞ and
{fn } is a bounded sequence in Lp (X), then there exists a subsequence {fn k }
such that fn k * f in Lp (X) weakly.

Definition 2.40. A sequence {fn } ⊂ L∞ (X) is said to converge weak-* to


f ∈ L∞ (X) if
Z Z
fn (x)φ(x)dx → f (x)φ(x)dx ∀φ ∈ L1 (X) .
X X

We denote this convergence by saying that fn * f in L∞ (X) weak-*.

Theorem 2.41 (Weak-* compactness for L∞ ). If {fn } is a bounded se-



quence in L∞ (X), then there exists a subsequence {fn k } such that fn k * f
in L∞ (X) weak-*.
2.1. LP SPACES 27

Lemma 2.42. If fn → f in Lp (X), then fn * f in Lp (X).

Proof. By Hölder’s inequality,


Z

g(fn − f )dx ≤ kfn − f kLp kgkLq .

X

Note that if fn is weakly convergent, in general, this does not imply that
fn is strongly convergent.

Example 2.43. If p = 2, let fn denote any orthonormal sequence in L2 (X).


From Bessel’s inequality
∞ Z
X

fn gdx ≤ kgk2 2

L (X) ,
n=1 X

we see that fn * 0 in L2 (X).

This example shows that the map f 7→ kf kLp is continuous, but not
weakly continuous. It is, however, weakly lower-semicontinuous.

Theorem 2.44. If fn * f weakly in Lp (X), then kf kLp ≤ lim inf n→∞ kfn kLp .

Proof. As a consequence of Theorem 2.31,


Z Z

kf kLp (X) = sup f gdx = sup lim fn gdx
kgkLq (X) =1 X kgkLq (X) =1 n→∞ X

≤ sup lim inf kfn kLp kgkLq .


kgkLq (X) =1 n→∞

Theorem 2.45. If fn * f in Lp (X), then fn is bounded in Lp (X).

Theorem 2.46. Suppose that Ω ⊂ Rn is a bounded. Suppose that

sup kfn kLp (Ω) ≤ M < ∞ and fn → f a.e.


n

If 1 < p < ∞, then fn * f in Lp (Ω).


28 CHAPTER 2. INTRODUCTION TO SOBOLEV SPACES

Proof. Egoroff’s theorem states that for all  > 0, there exists E ⊂ Ω such
that µ(E) <  and fn → f uniformly on E c . By definition, fn * f in Lp (Ω)
p
for p ∈ (1, ∞) if Ω (fn − f )gdx → 0 for all g ∈ Lq (Ω), q = p−1
R
. We have the
inequality
Z Z Z
(fn − f )gdx ≤ |fn − f | |g| dx + |fn − f | |g| dx .
Ω E Ec

Choose n ∈ N sufficiently large, so that |fn (x) − f (x)| ≤ δ for all x ∈ E c .


By Hölder’s inequality,
Z
|fn − f | |g| dx ≤ kfn − f kLp (E c ) kgkLq (E c ) ≤ δµ(E c )kgkLq (Ω) ≤ Cδ
Ec

for a constant C < ∞.


By the Dominated Convergence Theorem, kfn − f kLp (Ω) ≤ 2M so by
Hölder’s inequality, the integral over E is bounded by 2M kgkLq (E) . Next, we
use the fact that the integral is continuous with respect to the measure of the
set over which the integral is taken. In particular, if 0 ≤ h is integrable, then
for all δ > 0, there exists  > 0 such that if the set E has measure µ(E ) < ,
R
then E hdx ≤ δ. To see this, either approximate h by simple functions, or
R
use the Dominated Convergence theorem for the integral Ω 1E (x)h(x)dx.


Remark 2.47. The proof of Theorem 2.46 does not work in the case that
p = 1, as Hölder’s inequality gives
Z
|fn − f | |g| dx ≤ kfn − f kL1 (Ω) kgkL∞ (E) ,
E

so we lose the smallness of the right-hand side.

Remark 2.48. Suppose that E ⊂ X is bounded and measurable, and let


g = 1E . If fn * f in Lp (X), then
Z Z
fn (x)dx → f (x)dx;
E E

hence, if fn * f , then the average of fn converges to the average of f


pointwise.
2.1. LP SPACES 29

2.1.13 Integral operators

If u : Rn → R satisfies certain integrability conditions, then we can define


the operator K acting on the function u as follows:
Z
Ku(x) = k(x, y)u(y)dy ,
Rn
where k(x, y) is called the integral kernel.The mollification procedure, intro-
duced in Definition 2.27, is one example of the use of integral operators; the
Fourier transform is another.

Definition 2.49. Let L(Lp (Rn ), Lp (Rn )) denote the space of bounded linear
operators from Lp (Rn ) to itself. Using the Representation Theorem 2.31, the
natural norm on L(Lp (Rn ), Lp (Rn )) is given by
Z

kKkL(Lp (Rn ),Lp (Rn )) = sup sup Kf (x)g(x)dx .
kf kLp =1 kgkLq =1 n
R
R
Theorem 2.50. Let 1 ≤ p < ∞, Ku(x) = Rn k(x, y)u(y)dy, and suppose
that
Z Z
n
|k(x, y)|dx ≤ C1 ∀y ∈ R and |k(x, y)|dy ≤ C2 ∀x ∈ Rn ,
Rn Rn
where 0 < C1 , C2 < ∞. Then K : Lp (Rn ) → Lp (Rn ) is bounded and
1 p−1
kKkL(Lp (Rn ),Lp (Rn )) ≤ C1p C2 p .

In order to prove Theorem 2.50, we will need another well-known in-


equality.

Lemma 2.51 (Cauchy-Young Inequality). If p1 + 1q = 1, then for all a, b ≥ 0,


ap bq
ab ≤ + .
p q
Proof. Suppose that a, b > 0, otherwise the inequality trivially holds.

ab = exp(log(ab)) = exp(log a + log b) (since a, b > 0)


 
1 p 1 q
= exp log a + log b
p q
1 1
≤ exp(log ap ) + exp(log bq ) (using the convexity of exp)
p q
ap bq
= +
p q
1 1
where we have used the condition p + q = 1. 
30 CHAPTER 2. INTRODUCTION TO SOBOLEV SPACES

1
Lemma 2.52 (Cauchy-Young Inequality with δ). If p + 1q = 1, then for all
a, b ≥ 0,
ab ≤ δ ap + Cδ bq , δ > 0,

with Cδ = (δp)−q/p q −1 .

Proof. This is a trivial consequence of Lemma 2.51 by setting


b
ab = a · (δp)1/p .
(δp)1/p

|f (y)|p
Proof of Theorem 2.50. According to Lemma 2.51, |f (y)g(x)| ≤ p +
|g(x)|q
q so that
Z Z

n n k(x, y)f (y)g(x)dydx

R R
|k(x, y)| |k(x, y)|
Z Z Z Z
p
≤ dx|f (y)| dy + dy|g(x)|q dx
Rn Rn p Rn Rn q
C1 C2
≤ kf kpLp + kgkqLq .
p q
To improve this bound, notice that
Z Z

n n k(x, y)f (y)g(x)dydx

R R
|k(x, y)| |k(x, y)|
Z Z Z Z
≤ p
dx|tf (y)| dy + dy|t−1 g(x)|q dx
Rn Rn p Rn Rn q
C1 tp C2 t−q
≤ kf kpLp + kgkqLq =: F (t) .
p q
Find the value of t for which F (t) has a minimum to establish the desired
bounded. 

Theorem 2.53 (Simple version of Young’s inequality). Suppose that k ∈


L1 (Rn ) and f ∈ Lp (Rn ). Then

kk ∗ f kLp ≤ kkkL1 kf kLp .

Proof. Define Z
Kk (f ) = k ∗ f := k(x − y)f (y)dy .
Rn
Let C1 = C2 = kkkL1 (Rn ) . Then according to Theorem 2.50, Kk : Lp (Rn ) →
Lp (Rn ) and kKk kL(Lp (Rn ),Lp (Rn )) ≤ C1 . 
2.1. LP SPACES 31

Theorem 2.50 can easily be generalized to the setting of integral op-


erators K : Lq (Rn ) → Lr (Rn ) built with kernels k ∈ Lp (Rn ) such that
1 1
1+ r = p + 1q . Such a generalization leads to

Theorem 2.54 (Young’s inequality). Suppose that k ∈ Lp (Rn ) and f ∈


Lq (Rn ). Then
1 1 1
kk ∗ f kLr ≤ kkkLp kf kLq for 1 + = + .
r p q

2.1.14 Appendix 1: The monotone and dominated conver-


gence theorems and Fatou’s lemma

Let Ω ⊂ Rd denote an open and smooth subset. The domain Ω is called


smooth whenever its boundary ∂Ω is a smooth (d − 1)-dimensional hyper-
surface.

Theorem 2.55 (Monotone Convergence Theorem). Let fn : Ω → R∪{+∞}


denote a sequence of functions, fn ≥ 0, and suppose that the sequence fn is
monotonically increasing, i.e.,

f1 ≤ f2 ≤ f3 ≤ · · ·

Then Z Z
lim fn (x)dx = lim fn (x)dx .
n→∞ Ω Ω n→∞

Lemma 2.56 (Fatou’s Lemma). Suppose the sequence fn : Ω → R and


fn ≥ 0. Then
Z Z
lim inf fn (x)dx ≤ lim inf fn (x)dx .
Ω n→∞ n→∞ Ω

Example 2.57. Consider Ω = (0, 1) ⊂ R and suppose that fn = n1(0,1/n) .


R1 R1
Then 0 fn (x)dx = 1 for all n ∈ N, but lim inf n→∞ 0 fn (x)dx = 0.

Theorem 2.58 (Dominated Convergence Theorem). Suppose the sequence


fn : Ω → R, fn (x) → f (x) almost everywhere (with respect to Lebesgue
measure), and furthermore, |fn | ≤ g ∈ L1 (Ω). Then f ∈ L1 (Ω) and
Z Z
lim fn (x)dx = f (x)dx .
n→∞ Ω Ω

Equivalently, fn → f in L1 (Ω) so that limn→∞ kfn − f kL1 (Ω) = 0.


32 CHAPTER 2. INTRODUCTION TO SOBOLEV SPACES

In the exercises, you will be asked to prove that the Monotone Conver-
gence Theorem implies Fatou’s Lemma which, in turn, implies the Domi-
nated Convergence Theorem.

2.1.15 Appendix 2: The Fubini and Tonelli Theorems

Let (X, A, µ) and (Y, B, ν) denote two fixed measure spaces. The product
σ-algebra A × B of subsets of X × Y is defined by

A × B = {A × B : A ∈ A, B ∈ B}.

The set function µ × ν : A × B → [0, ∞] defined by

(µ × ν)(A × B) = µ(A) · ν(B)

for each A × B ∈ A × B is a measure.

Theorem 2.59 (Fubini). Let f : X × Y → R be a µ × ν-integrable function.


Then both iterated integrals exist and
Z Z Z Z Z
f d(µ × ν) = f dµdν = f dνdµ .
X×Y Y X X Y

The existence of the iterated integrals is by no means enough to ensure


that the function is integrable over the product space. As an example, let
X = Y = [0, 1] and µ = ν = λ with λ the Lebesgue measure. Set
(
x2 −y 2
(x2 +y 2 )2
,(x, y) 6= (0, 0)
f (x, y) = .
0, (x, y) = (0, 0)

Then a standard computation shows that


1Z 1 1Z 1
π π
Z Z
f (x, y)dxdy = − , f (x, y)dydx = .
0 0 4 0 0 4

Fubini’s theorem shows, of course, that f is not integrable over [0, 1]2
There is a converse to Fubini’s theorem, however, according to which
the existence of one of the iterated integrals is sufficient for the integrability
of the function over the product space. The theorem is known as Tonelli’s
theorem, and this result is often used.
2.1. LP SPACES 33

Theorem 2.60 (Tonelli). Let (X, A, µ) and (Y, B, ν) denote two σ-finite
measure spaces, and let f : X × Y → R be a µ × ν-measurable function. If
R R R R
one of the iterated integrals X Y |f |dνdµ or Y X |f |dµdν exists, then the
function f is µ × ν-integrable and hence, the other iterated integral exists
and Z Z Z Z Z
f d(µ × ν) = f dµdν = f dνdµ .
X×Y Y X X Y

2.1.16 Exercises

Problem 2.1. Use the Monotone Convergence Theorem to prove Fatou’s


Lemma.

Problem 2.2. Use Fatou’s Lemma to prove the Dominated Convergence


Theorem.

Problem 2.3. Let Ω ⊂ Rd denote an open and smooth subset. Let (a, b) ⊂ R
be an open interval, and let f : (a, b) × Ω → R be a function such that for
df
each t ∈ (a, b), f (t, ·) : Ω → R is integrable and dt (t, x) exists for each
(t, x) ∈ (a, b) × Ω. Futhermore, assume that there is an integrable function
g : Ω → [0, ∞) such that supt∈(a,b) | df (t, x)| ≤ g(x) for all x ∈ Ω. Show that
R dt
the function h defined by h(t) ≡ Ω f (t, x)dx is differentiable and that the
derivative is given by
dh d df
Z Z
(t) = f (t, x)dx = (t, x)dx
dt dt Ω Ω dt
for each t ∈ (a, b). Hint: You will need to use the definition of the deriva-
dr
tive for a real valued function function r : (a, b) → R which is dt (t0 ) =
limh→0 r(t0 +h)−r(t
h
0)
, as well as the Mean Value Theorem from calculus which
states the following: Let (t1 , t2 ) ⊂ R and let q : (t1 , t2 ) → R be differentiable
|q(t2 )−q(t1 )| dq 0
on (t1 , t2 ). Then t2 −t1 = dt (t ) where t1 is some point between t1 and
t2 .

Problem 2.4. Let Ω denote an open subset of Rn . If f ∈ L1 (Ω) ∩ L∞ (Ω),


show that f ∈ Lp (Ω) for 1 < p < ∞. If Ω is bounded, then show that
limp%∞ kf kLp = kf kL∞ . (Hint: For  > 0, you can prove that the set
E = {x ∈ Ω : |f (x)| > kf kL∞ − } has positive Lebesgue measure, and the
inequality [kf kL∞ − ] 1E ≤ |f | holds.)
34 CHAPTER 2. INTRODUCTION TO SOBOLEV SPACES

Problem 2.5. Theorem 2.17 states that if 1 ≤ p < ∞, f ∈ Lp , {fn } ⊂ Lp ,


fn → f a.e., and limn→∞ kfn kLp = kf kLp , then limn→∞ kfn − f kLp → 0.
Show by an example that this theorem is false when p = ∞.

Problem 2.6. Show that equality holds in the inequality

aλ b1−λ ≤ λa + (1 − λ)b, λ ∈ (0, 1), a, b ≥ 0

if and only if a = b. Use this to show that if f ∈ Lp and g ∈ Lq for


1 1
1 < p, q < ∞ and p + q = 1, then
Z
|f g|dx = kf kLp kgkLq

holds if and only if there exists two constants C1 and C2 (not both zero) such
that C1 |f |p = C2 |g|q holds.

Problem 2.7. Use the result of Problem 2.6 to prove that if f, g ∈ L3 (Ω)
satisfy Z
kf kL3 = kgkL3 = f 2 g dx = 1 ,

then g = |f | a.e.

Problem 2.8. Given f ∈ L1 (S1 ), 0 < r < 1, define


∞ 2π
1
Z
fˆn r|n| einθ , fˆn =
X
Pr f (θ) = f (θ)e−inθ dθ .
n=−∞
2π 0

Show that

1
Z
Pr f (θ) = pr ∗ f (θ) = pr (θ − φ)f (φ)dφ ,
2π 0

where

X 1 − r2
pr (θ) = r|n| einθ = .
n=−∞
1 − 2r cos θ + r2
1
R 2π
Show that 2π 0 pr (θ)dθ = 1.
2.1. LP SPACES 35

Problem 2.9. If f ∈ Lp (S1 ), 1 ≤ p < ∞, show that

Pr f → f in Lp (S1 ) as r % 1 .

Problem 2.10. Suppose that Y = [0, 1]2 is the unit square in R2 and let
a(y) denote a Y -periodic function in L∞ (R2 ). For  > 0, let a (x) = a( x ),
R ∗
and let ā = Y a(y)dy denote the average value of a. Prove that a * ā as
 → 0.


Problem 2.11. Let fn = n1(0, 1 ) . Prove that fn * 0 in L2 (0, 1), that
n
fn → 0 in L1 (0, 1), but that fn does not converge strongly in L2 (0, 1).
36 CHAPTER 2. INTRODUCTION TO SOBOLEV SPACES

2.2 The Sobolev spaces H k (Ω) for integers k ≥ 0


2.2.1 Weak derivatives

Definition 2.61 (Test functions). For Ω ⊂ Rn , set

C0∞ (Ω) = {u ∈ C ∞ (Ω) | spt(u) ⊂ V ⊂⊂ Ω},

the smooth functions with compact support. Traditionally D(Ω) is often used
to denote C0∞ (Ω), and D(Ω) is often referred to as the space of test functions.

du
For u ∈ C 1 (R), we can define dx by the integration-by-parts formula;
namely,

du dφ
Z Z
(x)φ(x)dx = − u(x) (x)dx ∀φ ∈ C0∞ (R) .
R dx R dx

Notice, however, that the right-hand side is well-defined, whenever u ∈


L1loc (R)

Definition 2.62. An element α ∈ Zn+ (nonnegative integers) is called a


∂ α1 ∂ αn
multi-index. For such an α = (α1 , ..., αn ), we write Dα = α
∂x1 1
··· ∂xα n and
n

|α| = α1 + · · ·αn .

Example 2.63. Let n = 2. If |α| = 0, then α = (0, 0); if |α| = 1, then


α = (1, 0) or α = (0, 1). If |α| = 2, then α = (1, 1).

Definition 2.64 (Weak derivative). Suppose that u ∈ L1loc (Ω). Then v α ∈


L1loc (Ω) is called the αth weak derivative of u, written v α = Dα u, if
Z Z
|α|
α
u(x) D φ(x)dx = (−1) v α (x)φ(x)dx ∀φ ∈ C0∞ (Ω) .
Ω Ω

Example 2.65. Let n = 1 and set Ω = (0, 2). Define the function

x, 0 ≤ x < 1
u(x) = .
1, 1 ≤ x ≤ 2

Then the function



1, 0 ≤ x < 1
v(x) =
0, 1 ≤ x ≤ 2
2.2. THE SOBOLEV SPACES H K (Ω) FOR INTEGERS K ≥ 0 37

is the weak derivative of u. To see this, note that for φ ∈ C0∞ (0, 2),
2 1 2
dφ dφ dφ
Z Z Z
u(x) (x)dx = x (x)dx + (x)dx
0 dx 0 dx 1 dx
Z 1 Z 1
=− φ(x)dx + xφ|10 + φ|21 =− φ(x)dx
0 0
Z 2
=− v(x)φ(x)dx .
0

Example 2.66. Let n = 1 and set Ω = (0, 2). Define the function

x, 0 ≤ x < 1
u(x) = .
2, 1 ≤ x ≤ 2

Then the weak derivative does not exist!


To prove this, assume for the sake of contradiction that there exists v ∈
L1loc (Ω) such that for all φ ∈ C0∞ (0, 2),
2 2

Z Z
v(x)φ(x)dx = − u(x) (x)dx .
0 0 dx
Then
2 Z 1 Z 2
dφ dφ
Z
v(x)φ(x)dx = − x (x)dx − 2 (x)dx
0 0 dx 1 dx
Z 1
= φ(x)dx − φ(1) + 2φ(1)
0
Z 1
= φ(x)dx + φ(1) .
0

Suppose that φj is a sequence in C0∞ (0, 2) such that φj (1) = 1 and φj (x) → 0
for x 6= 1. Then
Z 2 Z 1
1 = φj (1) = v(x)φj (x)dx − φj (x)dx → 0 ,
0 0

which provides the contradiction.

Definition 2.67. For p ∈ [1, ∞], define W 1,p (Ω) = {u ∈ Lp (Ω) | weak derivative exists , Du ∈
Lp (Ω)}, where Du is the weak derivative of u.

Example 2.68. Let n = 1 and set Ω = (0, 1). Define the function f (x) =
du
sin(1/x). Then u ∈ L1 (0, 1) and dx = − cos(1/x)/x2 ∈ L1loc (0, 1), but u 6∈
W 1,p (Ω) for any p.
38 CHAPTER 2. INTRODUCTION TO SOBOLEV SPACES

Definition 2.69. In the case p = 2, we set H 1 (Ω) = W 1,p (Ω).

Example 2.70. Let Ω = B(0, 1) ⊂ R2 and set u(x) = |x|−α . We want to


determine the values of α for which u ∈ H 1 (Ω).
Since |x|−α = 3j=1 (xj xj )−α/2 , then ∂xi |x|−α = −α|x|−α−2 xi is well-
P

defined away from x = 0.

Step 1. We show that u ∈ L1loc (Ω). To see this, note that −α dx


R
Ω |x| =
R 2π R 1 −α
0 0 r rdrdθ < ∞ whenever α < 2.

Step 2. Set the vector v(x) = −α|x|−α−2 x (so that each component is given
by vi (x) = −α|x|−α−2 xi ). We show that
Z Z
u(x)Dφ(x)dx = − v(x)φ(x)dx ∀φ ∈ C0∞ (B(0, 1)) .
B(0,1) B(0,1)

To see this, let Ωδ = B(0, 1) − B(0, δ), let n denote the unit normal to ∂Ωδ
(pointing toward the origin). Integration by parts yields
Z Z 2π Z
−α −α
|x| Dφ(x)dx = δ φ(x)n(x)δdθ + α |x|−α−2 x φ(x)dx .
Ωδ 0 Ωδ

R 2π
Since limδ→0 δ 1−α 0 φ(x)n(x)dθ = 0 if α < 1, we see that
Z Z
−α
lim |x| Dφ(x)dx = lim α |x|−α−2 x φ(x)dx
δ→0 Ωδ δ→0 Ωδ

R 2π R 1
Since 0 0 r−α−1 rdrdθ < ∞ if α < 1, the Dominated Convergence Theo-
rem shows that v is the weak derivative of u.
R 2π R 1
Step 3. v ∈ L2 (Ω), whenever 0 0 r−2α−2 rdrdθ < ∞ which holds if α < 0.

Remark 2.71. Note that if the weak derivative exists, it is unique. To see
this, suppose that both v1 and v2 are the weak derivative of u on Ω. Then

R
Ω (v1 − v2 )φdx = 0 for all φ ∈ C0 (Ω), so that v1 = v2 a.e.

2.2.2 Definition of Sobolev Spaces

Definition 2.72. For integers k ≥ 0 and 1 ≤ p ≤ ∞,

W k,p (Ω) = {u ∈ L1loc (Ω) |Dα u exists and is in Lp (Ω) for |α| ≤ k}.
2.2. THE SOBOLEV SPACES H K (Ω) FOR INTEGERS K ≥ 0 39

Definition 2.73. For u ∈ W k,p (Ω) define


 1
p

ukpLp (Ω) 
X
α
kukW k,p (Ω) =  kD for 1 ≤ p < ∞ ,
|α|≤k

and
X
kukW k,∞ (Ω) = kDα ukL∞ (Ω) .
|α|≤k

The function k · kW k,p (Ω) is clearly a norm since it is a finite sum of Lp


norms.

Definition 2.74. A sequence uj → u in W k,p (Ω) if limj→∞ kuj −ukW k,p (Ω) =
0.

Theorem 2.75. W k,p (Ω) is a Banach space.

Proof. Let uj denote a Cauchy sequence in W k,p (Ω). It follows that for all
|α| ≤ k, Dα uj is a Cauchy sequence in Lp (Ω). Since Lp (Ω) is a Banach
space (see Theorem 2.19), for each α there exists uα ∈ Lp (Ω) such that

Dα uj → uα in Lp (Ω) .

When α = (0, ..., 0) we set u := u(0,...,0) so that uj → u in Lp (Ω). We must


show that uα = Dα u.
For each φ ∈ C0∞ (Ω),
Z Z
α
uD φdx = lim uj Dα φdx
Ω j→∞ Ω
Z
|α|
= (−1) lim Dα uj φdx
j→∞ Ω
Z
= (−1)|α| uα φdx ;

thus, uα = Dα u and hence Dα uj → Dα u in Lp (Ω) for each |α| ≤ k, which


shows that uj → u in W k,p (Ω). 

Definition 2.76. For integers k ≥ 0 and p = 2, we define

H k (Ω) = W k,2 (Ω) .

H k (Ω) is a Hilbert space with inner-product (u, v)H k (Ω) = α u, D α v)


P
|α|≤k (D L2 (Ω) .
40 CHAPTER 2. INTRODUCTION TO SOBOLEV SPACES

2.2.3 A simple version of the Sobolev embedding theorem

For two Banach spaces B1 and B2 , we say that B1 is embedded in B2 if


kukB2 ≤ CkukB1 for some constant C and for u ∈ B1 . We wish to determine
which Sobolev spaces W k,p (Ω) can be embedded in the space of continuous
functions. To motivate the type of analysis that is to be employed, we study
a special case.

Theorem 2.77 (Sobolev embedding in 2-D). For kp ≥ 2,

max |u(x)| ≤ CkukW k,p (R2 ) ∀u ∈ C0∞ (Ω) . (2.7)


x∈R2

Proof. Given u ∈ C0∞ (Ω), we prove that for all x ∈ spt(u),

|u(x)| ≤ CkDα u(x)kLp (Ω) ∀|α| ≤ k .

By choosing a coordinate system centered about x, we can assume that


x = 0; thus, it suffices to prove that

|u(0)| ≤ CkDα u(x)kLp (Ω) ∀|α| ≤ k .

Let g ∈ C ∞ ([0, ∞)) with 0 ≤ g ≤ 1, such that g(x) = 1 for x ∈ [0, 21 ] and
g(x) = 0 for x ∈ [ 34 , ∞).
By the fundamental theorem of calculus,
Z 1 Z 1
u(0) = − ∂r [g(r)u(r, θ)]dr = − ∂r (r) ∂r [g(r)u(r, θ)]dr
0 0
Z 1
= r ∂r2 [g(r)u(r, θ)]dr
0
Z 1 Z 1
(−1)k k−1 k (−1)k
= r ∂r [g(r)u(r, θ)]dr = rk−2 ∂rk [g(r)u(r, θ)]rdr
(k − 1)! 0 (k − 1)! 0

Integrating both sides from 0 to 2π, we see that


2π 1
(−1)k
Z Z
u(0) = rk−2 ∂rk [g(r)u(r, θ)]rdrdθ .
2π(k − 1)! 0 0

The change of variables from Cartesian to polar coordinates is given by

x(r, θ) = r cos θ , y(r, θ) = r sin θ .


2.2. THE SOBOLEV SPACES H K (Ω) FOR INTEGERS K ≥ 0 41

By the chain-rule,

∂r u(x(r, θ), y(r, θ)) = ∂x u cos θ + ∂y u sin θ ,


∂r2 u(x(r, θ), y(r, θ)) = ∂x2 u cos2 θ + 2∂xy
2
u cos θ sin θ + ∂y2 u sin2 θ
..
.

It follows that ∂rk = aα (θ)Dα , where aα consists of trigonometric


P
|α|≤k
polynomials of θ, so that

(−1)k
Z X
u(0) = rk−2 aα (θ)Dα [g(r)u(x)]dx
2π(k − 1)! B(0,1) |α|≤k
X
≤ Ckrk−2 kLq (B(0,1)) kDα (gu)kLp (B(0,1))
|α|≤k
Z 1  p−1
p
p(k−2)
≤C r p−1 rdr kukW k,p (R2 ) .
0

p(k−2)
Hence, we require p−1 + 1 > −1 or kp > 2. 

2.2.4 Approximation of W k,p (Ω) by smooth functions

Recall that Ω = {x ∈ Ω | dist(x, ∂Ω) > }.

Theorem 2.78. For integers k ≥ 0 and 1 ≤ p < ∞, let

u = η ∗ u in Ω ,

where η is the standard mollifier defined in Definition 2.25. Then

(A) u ∈ C ∞ (Ω ) for each  > 0, and

k,p
(B) u → u in Wloc (Ω) as  → 0.

k,p
Definition 2.79. A sequence uj → u in Wloc (Ω) if uj → u in W k,p (Ω̃) for
each Ω̃ ⊂⊂ Ω.

Proof of Theorem 2.78. Theorem 2.28 proves part (A). Next, let v α denote
the the αth weak partial derivative of u. To prove part (B), we show that
42 CHAPTER 2. INTRODUCTION TO SOBOLEV SPACES

Dα u = η ∗ v α in Ω . For x ∈ Ω ,
Z
α  α
D u (x) = D η (x − y)u(y)dy
Z Ω

= Dxα η (x − y)u(y)dy
Ω Z
|α|
= (−1) Dyα η (x − y)u(y)dy
Z Ω

= η (x − y)v α (y)dy = (η ∗ v α )(x) .


By part (D) of Theorem 2.28, Dα u → v α in Lploc (Ω). 

It is possible to refine the above interior approximation result all the


way to the boundary of Ω. We record the following theorem without proof.

Theorem 2.80. Suppose that Ω ⊂ Rn is a smooth, open, bounded subset,


and that u ∈ W k,p (Ω) for some 1 ≤ p < ∞ and integers k ≥ 0. Then there
exists a sequence uj ∈ C ∞ (Ω) such that

uj → u in W k,p (Ω) .

It follows that the inequality (2.7) holds for all u ∈ W k,p (R2 ).

2.2.5 Hölder Spaces

Recall that for Ω ⊂ Rn open and smooth, the class of Lipschitz functions
u : Ω → R satisfies the estimate

|u(x) − u(y)| ≤ C|x − y| ∀x, y ∈ Ω

for some constant C.

Definition 2.81 (Classical derivative). A function u : Ω → R is differen-


tiable at x ∈ Ω if there exists f : Ω → L(Rn ; Rn ) such that

|u(x) − u(y) − f (x) · (x − y)|


→ 0.
|x − y|

We call f (x) the classical derivative (or gradient) of u(x), and denote it by
Du(x).
2.2. THE SOBOLEV SPACES H K (Ω) FOR INTEGERS K ≥ 0 43

Definition 2.82. If u : Ω → R is bounded and continuous, then

kukC 0 (Ω) = max |u(x)| .


x∈Ω

If in addition u has a continuous and bounded derivative, then

kukC 1 (Ω) = kukC 0 (Ω) + kDukC 0 (Ω) .

The Hölder spaces interpolate between C 0 (Ω) and C 1 (Ω).

Definition 2.83. For 0 < γ ≤ 1, the space C 0,γ (Ω) consists of those func-
tions for which

kukC 0,γ (Ω) := kukC 0 (Ω) + [u]C 0,γ (Ω) < ∞ ,

where the γth Hölder semi-norm [u]C 0,γ (Ω) is defined as

|u(x) − u(y)|
 
[u]C 0,γ (Ω) = max .
x,y∈Ω |x − y|γ
x6=y

The space C 0,γ (Ω) is a Banach space.

2.2.6 Morrey’s inequality

We can now offer a refinement and extension of the simple version of the
Sobolev Embedding Theorem 2.77.

Theorem 2.84 (Morrey’s inequality). For n < p ≤ ∞, let B(x, r) ⊂ Rn


and let y ∈ B(x, r). Then

1− n
|u(x) − u(y)| ≤ Cr p kDukLp (B(x,2r)) ∀u ∈ C 1 (Rn ) .

In fact, Morrey’s inequality holds for all u ∈ W 1,p (B(x, 2r)) (see Problem
2.25 in the Exercises).

Notation 2.85 (Averaging). Let B(0, 1) ⊂ Rn . The volume of B(0, 1) is


n
π2
given by αn = Γ( n +1) and the surface area is |Sn−1 | = nαn . We define
2

1
Z Z
− f (y)dy = f (y)dy
B(x,r) αn rn B(x,r)
1
Z Z
− f (y)dS = f (y)dS .
∂B(x,r) nαn rn−1 ∂B(x,r)
44 CHAPTER 2. INTRODUCTION TO SOBOLEV SPACES

Lemma 2.86. For B(x, r) ⊂ Rn , y ∈ B(x, r) and u ∈ C 1 (B(x, r)),

|Du(y)|
Z Z
− |u(y) − u(x)|dy ≤ C n−1
dy .
B(x,r) B(x,r) |x − y|

Proof. For some 0 < s < r, let y = x + sω where ω ∈ Sn−1 = ∂B(0, 1). By
the fundamental theorem of calculus, for 0 < s < r,
Z s
d
u(x + sω) − u(x) = u(x + tω)dt
dt
Z0 s
= Du(x + tω) ωdt .
0

Since |ω| = 1, it follows that


Z s
|u(x + sω) − u(x)| ≤ |Du(x + tω)|dt .
0

Thus, integrating over Sn−1 yields


Z Z sZ
|u(x + sω) − u(x)|dω ≤ |Du(x + tω)|dωdt
Sn−1 0 Sn−1
Z sZ
tn−1
≤ |Du(x + tω)| n−1 dωdt
n−1 t
Z0 S
|Du(y)|
≤ n−1
dy ,
B(x,r) |x − y|

where we have set y = x + tω.


Multipling the above inequality by sn−1 and integrating s from 0 to r
shows that
Z rZ
rn |Du(y)|
Z
n−1
|u(x + sω) − u(x)|dωs ds ≤ dy
0 Sn−1 n B(x,r) |x − y|n−1
|Du(y)|
Z
n
≤ Cαn r dy ,
B(x,r) |x − y|n−1

which proves the lemma. 

Proof of Theorem 2.84. Assume first that u ∈ C 1 (B(x, 2r)). Let D =


B(x, r) ∩ B(y, r) and set r = |x − y|. Then
Z
|u(x) − u(y)| = − |u(x) − u(y)|dz
ZD Z
≤ − |u(x) − u(z)|dz + − |u(y) − u(z)|dz .
D D
2.2. THE SOBOLEV SPACES H K (Ω) FOR INTEGERS K ≥ 0 45

Since D equals the intersection of two balls of radius r, it is clear that can
choose a constant C, depending only on the dimension n, such that

|D| |D|
= =C.
|B(x, r)| |B(y, r)|

It follows that
Z Z
|u(x) − u(y)| ≤ − |u(x) − u(z)|dz + − |u(y) − u(z)|dz
D D
Z 
C
Z
≤ |u(x) − u(z)|dz + |u(y) − u(z)|dz
|B(x, r)| D D
Z Z
≤ C− |u(x) − u(z)|dz + C− |u(y) − u(z)|dz .
B(x,r) B(y,r)

Thus, by Lemma 2.86,


Z Z Z
1−n
− |u(x)−u(z)|dz ≤ C |x−z| |Du(z)|dz ≤ C |x−z|1−n |Du(z)|dz
B(x,r) B(x,r) B(x,2r)

and
Z Z Z
1−n
− |u(x)−u(z)|dz ≤ C |x−z| |Du(z)|dz ≤ C |x−z|1−n |Du(z)|dz
B(y,r) B(y,r) B(x,2r)

so that
Z
|u(x) − u(y)| ≤ C |x − z|1−n |Du(z)|dz (2.8)
B(x,2r)

and by Hölder’s inequality,


! p−1 !1
Z p Z p
p(1−n)
n−1 p
|u(x) − u(y)| ≤ C s p−1 s dsdω |Du(z)| dz
B(0,2r) B(x,2r)

Morrey’s inequality implies the following embedding theorem.

Theorem 2.87 (Sobolev embedding theorem for k = 1). There exists a


constant C = C(p, n) such that

kuk 0,1− n ≤ CkukW 1,p (Rn ) ∀u ∈ W 1,p (Rn ) .


C p (Rn )
46 CHAPTER 2. INTRODUCTION TO SOBOLEV SPACES

Proof. First assume that u ∈ C01 (Rn ). Given Morrey’s inequality, it suffices
to show that max |u| ≤ CkukW 1,p (Rn ) . Using Lemma 2.86, for all x ∈ Rn ,
Z Z
|u(x)| ≤ − |u(x) − u(y)|dy + − |u(y)|dy
B(x,1) B(x,1)
|Du(y)|
Z
≤C dy + CkukLp (Rn )
B(x,1) |x − y|n−1
≤ CkukW 1,p (Rn ) ,

the last inequality following whenever p > n.


Thus,
kuk 0,1− n ≤ CkukW 1,p (Rn ) ∀u ∈ C 1 (Rn ) . (2.9)
C p (Rn )

By the density of C0∞ (Rn ) in W 1,p (Rn ), there is a sequence uj ∈ C0∞ (Rn )
such that
uj → u ∈ W 1,p (Rn ) .

By (2.9), for j, k ∈ N,

0,1− n
kuj − uk kC p (Rn ) ≤ Ckuj − uk kW 1,p (Rn ) .

0,1− n 0,1− n
Since C p (Rn ) is a Banach space, there exists a U ∈ C p (Rn ) such
that
0,1− n
uj → U in C p (Rn ) .

It follows that U = u a.e. in Ω. By the continuity of norms with respect to


strong convergence, we see that

kU k 0,1− n ≤ CkukW 1,p (Rn )


C p (Rn )

which completes the proof. 

In proving the above embedding theorem, we established that that for


p > n, we have the inequality

kukL∞ (Rn ) ≤ CkukW 1,p (Rn ) . (2.10)

We will see later that (2.10), via a scaling argument, leads to the following
important interpolation inequality: for p > n,
n p−n
kukL∞ (Rn ) ≤ C(n, p)kDukLp p (Rn ) kukLpp(Rn ) .
2.2. THE SOBOLEV SPACES H K (Ω) FOR INTEGERS K ≥ 0 47

Another important consequence of Morrey’s inequality is the relation-


ship between the weak and classical derivative of a function. We begin by
recalling the definition of classical differentiability. A function u : Rn → Rm
is differentiable at a point x if there exists a linear operator L : Rn → Rm
such that for each  > 0, there exists δ > 0 with |y − x| < δ implying that

ku(y) − u(x) − L(y − x)k ≤ ky − xk .

When such an L exists, we write Du(x) = L and call it the classical deriva-
tive.
As a consequence of Morrey’s inequality, we extract information about
the classical differentiability properties of weak derivatives.

Theorem 2.88 (Differentiability a.e.). If Ω ⊂ Rn , n < p ≤ ∞ and u ∈


1,p
Wloc (Ω), then u is differentiable a.e. in Ω, and its gradient equals its weak
gradient almost everywhere.

Proof. We first restrict n < p < ∞. By a version Lebesgue’s differentiation


theorem, for almost every x ∈ Ω,
Z
lim − |Du(x) − Du(z)|p dz = 0 , (2.11)
r→0 B(x,r)

where Du denotes the weak derivative of u. Thus, for r > 0 sufficiently


small, we see that
Z
− |Du(x) − Du(z)|p dz <  .
B(x,r)

Fix a point x ∈ Ω for which (2.11) holds, and define the function

wx (y) = u(y) − u(x) − Du(x) · (y − x) .

Notice that wx (x) = 0 and that

Dy wx (y) = Du(y) − Du(x) .

Set r = |x − y|. Since |u(y) − u(x) − Du(x) · (y − x)| = |wx (y) − wx (x)|, an
application of the inequality (2.8) that we obtained in the proof of Morrey’s
48 CHAPTER 2. INTRODUCTION TO SOBOLEV SPACES

inequality then yields the estimate

|Dz wx (z)|
Z
|u(y) − u(x) − Du(x) · (y − x)| ≤ C dz
B(x,2r) |x − z|n−1
|Du(z) − Du(x)|
Z
=C dz
B(x,2r) |x − z|n−1
!1
Z p
1− n
≤ Cr p |Du(z) − Du(x)|p dz
B(x,2r)
!1
Z p

≤ Cr − |Du(z) − Du(x)|p dz
B(x,2r)

≤ C|x − y| ,

from which it follows that Du(x) is the classical derivative of u at the point
x.
1,∞ 1,p
The case that p = ∞ follows from the inclusion Wloc (Ω) ⊂ Wloc (Ω) for
all 1 ≤ p < ∞. 

2.2.7 The Gagliardo-Nirenberg-Sobolev inequality

In the previous section, we considered the embedding for the case that p > n.

Theorem 2.89 (Gagliardo-Nirenberg inequality). For 1 ≤ p < n, set p∗ =


np
n−p . Then

kukLp∗ (Rn ) ≤ Cp,n kDukLp (Rn ) ∀u ∈ W 1,p (Rn ) .

Proof for the case n = 2. Suppose first that p = 1 in which case p∗ = 2, and
we must prove that

kukL2 (R2 ) ≤ CkDukL1 (R2 ) ∀u ∈ C01 (R2 ) . (2.12)

Since u has compact support, by the fundamental theorem of calculus,


Z x1 Z x2
u(x1 , x2 ) = ∂1 u(y1 , x2 )dy1 = ∂2 u(x1 , y2 )dy2
−∞ −∞

so that
Z ∞ Z ∞
|u(x1 , x2 )| ≤ |∂1 u(y1 , x2 )|dy1 ≤ |Du(y1 , x2 )|dy1
−∞ −∞
2.2. THE SOBOLEV SPACES H K (Ω) FOR INTEGERS K ≥ 0 49

and
Z ∞ Z ∞
|u(x1 , x2 )| ≤ |∂2 u(x1 , y2 )|dy2 ≤ |Du(x1 , y2 )|dy2 .
−∞ −∞

Hence, it follows that


Z ∞ Z ∞
2
|u(x1 , x2 )| ≤ |Du(y1 , x2 )|dy1 |Du(x1 , y2 )|dy2
−∞ −∞

Integrating over R2 , we find that


Z ∞Z ∞
|u(x1 , x2 )|2 dx1 dx2
−∞ −∞
Z ∞ Z ∞ Z ∞ Z ∞ 
≤ |Du(y1 , x2 )|dy1 |Du(x1 , y2 )|dy2 dx1 dx2
−∞ −∞ −∞ −∞
Z ∞ Z ∞ 2
≤ |Du(x1 , x2 )|dx1 dx2
−∞ −∞

which is (2.12).
Next, if 1 ≤ p < 2, substitute |u|γ for u in (2.12) to find that
Z 1 Z
2

|u| dx ≤ Cγ |u|γ−1 |Du|dx
R2 R2
Z  p−1
p(γ−1) p
≤ CγkDukLp (R2 ) |u| p−1 dx
R2

p(γ−1) p
Choose γ so that 2γ = p−1 ; hence, γ = 2−p , and

Z  2−p
2p 2p
|u| 2−p dx ≤ CγkDukLp (R2 ) ,
R2

so that
kuk 2p ≤ Cp,n kDukLp (Rn ) (2.13)
L 2−p (Rn )

for all u ∈ C01 (R2 ).


Since C0∞ (R2 ) is dense in W 1,p (R2 ), there exists a sequence uj ∈ C0∞ (R2 )
such that
uj → u in W 1,p (R2 ) .

Hence, by (2.13), for all j, k ∈ N,

kuj − uk k 2p ≤ Cp,n kDuj − Duk kLp (Rn )


L 2−p (Rn )
50 CHAPTER 2. INTRODUCTION TO SOBOLEV SPACES

2p
so there exists U ∈ L 2−p (Rn ) such that
2p
uj → U in L 2−p (Rn ) .

Hence U = u a.e. in R2 , and by continuity of the norms, (2.13) holds for all
u ∈ W 1,p (R2 ). 

Proof for the general case of dimension n. Following the proof for n = 2,
we see that
1
n
Z ∞ 
n−1
|u(x)| n−1 ≤ Πni=1 |Du(x1 , ..., yi , ..., xn )|dyi
−∞

so that
1
Z ∞ n
Z ∞ Z ∞ 
n−1
|u(x)| n−1 dx1 ≤ Πni=1 |Du(x1 , ..., yi , ..., xn )|dyi dx1
−∞ −∞ −∞
1 1
Z ∞ 
n−1
Z ∞ Z ∞ 
n−1
|Du|dy1 Πni=2 |Du|dyi dx1
−∞ −∞ −∞
1 1
Z ∞ 
n−1
Z ∞ Z ∞ 
n−1
|Du|dy1 Πni=2 |Du|dx1 dyi ,
−∞ −∞ −∞

where the last inequality follows from Hölder’s inequality.


Integrating the last inequality with respect to x2 , we find that
1
Z ∞ Z ∞ n
Z ∞ Z ∞ 
n−1
Z ∞ 1
|u(x)| n−1 dx1 dx2 < |Du|dx1 dy2 Πni=1 Iin−1 dx2 ,
−∞ −∞ −∞ −∞ −∞ i6=2

where
Z ∞ Z ∞ Z ∞
I1 = |Du|dy1 , Ii = |Du|dx1 dyi for i = 3, ..., n .
−∞ −∞ −∞

Applying Hölder’s inequality, we find that


Z ∞Z ∞
n
|u(x)| n−1 dx1 dx2
−∞ −∞
1 1
Z ∞ Z ∞ 
n−1
Z ∞ Z ∞ 
n−1
≤ |Du|dx1 dy2 |Du|dy1 dx2
−∞ −∞ −∞ −∞
1
Z ∞ Z ∞ Z ∞ 
n−1
Πni=3 |Du|dx1 dx2 dyi .
−∞ −∞ −∞
2.2. THE SOBOLEV SPACES H K (Ω) FOR INTEGERS K ≥ 0 51

Next, continue to integrate with respect to x3 , ..., xn to find that


1
Z
n
Z ∞ Z ∞ 
n−1
|u| n−1 dx ≤ Πni=1 ··· |Du|dx1 ...dyi ...dxn
Rn −∞ −∞
Z  n
n−1
= |Du|dx .
Rn

This proves the case that p = 1. The case that 1 < p < n follows identically
as in the proof of n = 2. 

It is common to employ the Gagliardo-Nirenberg inequality for the case


that p = 2; as stated, the inequality is not well-defined in dimension two,
but in fact, we have the following theorem.

Theorem 2.90. Suppose that u ∈ H 1 (R2 ). Then for all 1 ≤ q < ∞,


kukLq (R2 ) ≤ C qkukH 1 (R2 ) .

Proof. Let x and y be points in R2 , and write r = |x − y|. Let θ ∈ S1 .


Introduce spherical coordinates (r, θ) with origin at x, and let g be the
same cut-off function that was used in the proof of Theorem 2.77. Define
U := g(r)u(r, θ). Then

1 1
∂U ∂U
Z Z
u(x) = − (r, θ)dr − |x − y|−1 (r, θ)rdr
0 ∂r 0 ∂r

and
Z 1
|u(x)| ≤ |x − y|−1 |DU (r, θ)|rdr .
0

Integrating over S1 , we obtain:

1
Z
|u(x)| ≤ 1B(x,1) |x − y|−1 |DU (y)|dy := K ∗ |DU | ,
2π R2

1 −1
where the integral kernel K(x) = 2π 1B(0,1) |x| .
Using Young’s inequality from Theorem 2.54, we obtain the estimate

1 1 1
kK ∗ f kLq (R2 ) ≤ kKkLk (R2 ) kf kL2 (R2 ) for = − + 1. (2.14)
k q 2
52 CHAPTER 2. INTRODUCTION TO SOBOLEV SPACES

Using the inequality (2.14) with f = |DU |, we see that


"Z #1
k

kukLq (R2 ) ≤ CkDU kL2 (R2 ) |y|−k dy


B(0,1)
Z 1  k1
1−k
≤ CkDU kL2 (R2 ) r dr
0
 1
q+2 k
= CkukH 1 (R2 ) .
4
1
When q → ∞, k → 21 , so
1
kukLq (R2 ) ≤ Cq 2 kukH 1 (R2 ) .

Evidently, it is not possible to obtain the estimate kukL∞ (Rn ) ≤ CkukW 1,n (Rn )
with a constant C < ∞. The following provides an example of a function in
this borderline situation.

Example 2.91. Let Ω ⊂ R2 denote the open unit ball in R2 . The unbounded
 
1
function u = log log 1 + |x| belongs to H 1 (B(0, 1)).
First, note that
2π 1h
1 i2
Z Z Z 
|u(x)|2 dx = log log 1 + rdrdθ .
Ω 0 0 r

The only potential singularity of the integrand occurs at r = 0, but according


to L’Hospital’s rule,
h 1 i2 
lim r log log 1 + = 0, (2.15)
r→0 r

so the integrand is continuous and hence u ∈ L2 (Ω).


In order to compute the partial derivatives of u, note that
df
∂ xj d f (x) dz
|x| = , and |f (z)| = ,
∂xj |x| dz |f (z)|

where f : R → R is differentiable. It follows that for x away from the origin,

−x
Du(x) = 1 , (x 6= 0) .
log(1 + |x| )(|x| + 1)|x|2
2.2. THE SOBOLEV SPACES H K (Ω) FOR INTEGERS K ≥ 0 53

Let φ ∈ C0∞ (Ω) and fix  > 0. Then


∂φ ∂u
Z Z Z
u(x) (x)dx = − (x)φ(x)dx + uφNi dS ,
Ω−B (0) ∂xi Ω−B(0,) ∂xi ∂B(0,)

where N = (N1 , ..., Nn ) denotes the inward-pointing unit normal on the


curve ∂B(0, ), so that N dS = (cos θ, sin θ)dθ. It follows that
Z Z
u(x)Dφ(x)dx = − Du(x)φ(x)dx
Ω−B (0) Ω−B (0)

1
Z 
− (cos θ, sin θ) log log 1 + φ(, θ)dθ . (2.16)
0 

We claim that Du ∈ L2 (Ω) (and hence also in L1 (Ω)), for


2π 1
1
Z Z Z
2
|Du(x)| dx =  h i2 drdθ
Ω 0 0 r(r + 1)2 log 1 + 1r
1/2 Z 1
1 1
Z
≤ π 2
dr + π i2 dr
r(log r)
h 
0 1/2 r(r + 1)2 log 1 + 1
r

where we use the inequality log(1 + 1r ) ≥ log 1r = − log r ≥ 0 for 0 ≤ r ≤ 1.


The second integral on the right-hand side is clearly bounded, while
1/2 − log 2 − log 2
1 1 t 1
Z Z Z
dr = e dt = dx < ∞ ,
0 r(log r)2 −∞ t et
2
−∞ x2

so that Du ∈ L2 (Ω). Letting  → 0 in (2.16) and using (2.15) for the


boundary integral, by the Dominated Convergence Theorem, we conclude
that Z Z
u(x)Dφ(x)dx = − Du(x)φ(x)dx ∀φ ∈ C0∞ (Ω) .
Ω Ω

2.2.8 Local coordinates near ∂Ω

Let Ω ⊂ Rn denote an open, bounded subset with C 1 boundary, and let


{Ul }K
l=1 denote an open covering of ∂Ω, such that for each l ∈ {1, 2, ..., K},
with

Vl = B(0, rl ), denoting the open ball of radius rl centered at the origin and,
Vl+ = Vl ∩ {xn > 0} ,
Vl− = Vl ∩ {xn < 0} ,
54 CHAPTER 2. INTRODUCTION TO SOBOLEV SPACES

there exist C 1 -class charts θl which satisfy

θl : Vl → Ul is a C 1 diffeomorphism , (2.17)
θl (Vl+ ) = Ul ∩ Ω ,
θl (Vl ∩ {xn = 0}) = Ul ∩ ∂Ω .

2.2.9 Sobolev extensions and traces.

Let Ω ⊂ Rn denote an open, bounded domain with C 1 boundary.

Theorem 2.92. Suppose that Ω̃ ⊂ Rn is a bounded and open domain such


that Ω ⊂⊂ Ω̃. Then for 1 ≤ p ≤ ∞, there exists a bounded linear operator

E : W 1,p (Ω) → W 1,p (Rn )

such that for all u ∈ W 1,p (Ω),

1. Eu = u a.e. in Ω;

2. spt(Eu) ⊂ Ω̃;

3. kEukW 1,p (Rn ) ≤ CkukW 1,p (Ω) for a constant C = C(p, Ω, Ω̃).

Theorem 2.93. For 1 ≤ p < ∞, there exists a bounded linear operator

T : W 1,p (Ω) → Lp (Ω)

such that for all u ∈ W 1,p (Ω)

1. T u = u|∂Ω for all u ∈ W 1,p (Ω) ∪ C 0 (Ω);

2. kT ukLp (∂Ω) ≤ CkukW 1,p (Ω) for a constant C = C(p, Ω).

Proof. Suppose that u ∈ C 1 (Ω), z ∈ ∂Ω, and that ∂Ω is locally flat near
z. In particular, for r > 0 sufficiently small, B(z, r) ∪ ∂Ω ⊂ {xn = 0}. Let
0 ≤ ξ ∈ C0∞ (B(z, r)) such that ξ = 1 on B(z, r/2). Set Γ = ∂Ω ∪ B(z, r/2),
2.2. THE SOBOLEV SPACES H K (Ω) FOR INTEGERS K ≥ 0 55

B + (z, r) = B(z, r) ∪ Ω, and let dxh = dx1 · · · dxn−1 . Then


Z Z
p
|u| dxh ≤ ξ|u|p dxh
Γ {xn =0}

Z
=− (ξ|u|p )dx
+
B (z,r) ∂x n
∂ξ ∂u
Z Z
p
≤− |u| dx − p ξ|u|p−2 u dx
B + (z,r) ∂xn B + (z,2δ) ∂xn

∂u
Z
p p−1
≤C |u| dx + Ck|u| k p−1 p
B + (z,r) L (B + (z,r)) ∂xn p +
L (B (z,r))
Z
≤C (|u|p + |Du|p )dx . (2.18)
B + (z,r)

On the other hand, if the boundary is not locally flat near z ∈ ∂Ω,
then we use a C 1 diffeomorphism to locally straighten the boundary. More
specifically, suppose that z ∈ ∂Ω ∪ Ul for some l ∈ {1, ..., K} and consider
the C 1 chart θl defined in (2.17). Define the function U = u ◦ θl ; then
U : Vl+ → R. Setting Γ = Vl ∪ {xn = 0k, we see from the inequality (2.18),
that
Z Z
p
|U | dxh ≤ Cl (|U |p + |DU |p )dx .
Γ Vl+

Using the fact that Dθl is bounded and continuous on Vl+ , the change of
variables formula shows that
Z Z
|u|p dS ≤ Cl (|u|p + |Du|p )dx .
Ul ∪∂Ω Ul+

Summing over all l ∈ {1, ..., K} shows that


Z Z
|u| dS ≤ C (|u|p + |Du|p )dx .
p
(2.19)
∂Ω Ω

The inequality (2.19) holds for all u ∈ C 1 (Ω). According to Theorem 2.80,
for u ∈ W 1,p (Ω) there exists a sequence uj ∈ C ∞ (Ω) such that uj → u in
W 1,p (Ω). By inequality (2.19),

kT uk − T uj kLp (∂Ω) ≤ Ckuk − uj kW 1,p (Ω) ,

so that T uj is Cauchy in Lp (∂Ω), and hence a limit exists in Lp (∂Ω) We


define the trace operator T as this limit:

lim kT u − T uj kLp (∂Ω) = 0 .


j→0
56 CHAPTER 2. INTRODUCTION TO SOBOLEV SPACES

Since the sequence uj converges uniformly to u if u ∈ C 0 (Ω), we see that


T u = u|∂Ω for all u ∈ W 1,p (Ω) ∪ C 0 (Ω). 

Sketch of the proof of Theorem 2.92. Just as in the proof of the trace the-
orem, first suppose that u ∈ C 1 (ω) and that near z ∈ ∂Ω, ∂Ω is lo-
cally flat, so that for some r > 0, ∂Ω ∪ B(z, r) ⊂ {xn = 0}. Letting
B + = B(z, r) ∪ {xn ≥ 0} and B − = B(z, r) ∪ {xn ≤ 0} , we define the
extension of u by
if x ∈ B +

u(x)
ū(x) =
−3u(x1 , ..., xn−1 , −xn ) + 4u(x1 , ..., xn−1 , − x2n ) if x ∈ B − .
Define u+ = ū|B + and u− = ū|B − .
It is clear that u+ = u− on {xn = 0}, and by the chain-rule, it follows
that
∂u− ∂u− ∂u− xn
(x) = 3 (x1 , ..., −xn ) − 2 (x1 , ..., − ) ,
∂xn ∂xn ∂xn 2
∂u+ ∂u−
so that ∂xn = ∂xn on {xn = 0}. This shows that ū ∈ C 1 (B(z, r). using the
charts θl to locally straighten the boundary, and the density of the C ∞ (Ω)
in W 1,p (Ω), the theorem is proved. 

Later, we will provide a proof for higher-order Sobolev extensions of


H k -type functions.

2.2.10 The subspace W01,p (Ω)

Definition 2.94. We let W01,p (Ω) denote the closure of C0∞ (Ω) in W 1,p (Ω).

Theorem 2.95. Suppose that Ω ⊂ Rn is bounded with C 1 boundary, and


that u ∈ W 1,p (Ω). Then

u ∈ W01,p (Ω) iff T u = 0 on ∂Ω .

We can now state the Sobolev embedding theorems for bounded domains
Ω.

Theorem 2.96 (Gagliardo-Nirenberg inequality for W 1,p (Ω)). Suppose that


Ω ⊂ Rn is open and bounded with C 1 boundary, 1 ≤ p < n, and u ∈ W 1,p (Ω).
Then

kuk np ≤ CkukW 1,p (Ω) for a constant C = C(p, n, Ω) .


L n−p (Ω)
2.2. THE SOBOLEV SPACES H K (Ω) FOR INTEGERS K ≥ 0 57

Proof. Choose Ω̃ ⊂ Rn bounded such that Ω ⊂⊂ Ω̃, and let Eu denote the
Sobolev extension of u to Rn such that Eu = u a.e., spt(Eu) ⊂ Ω̃, and
kEukW 1,p (Rn ) ≤ CkukW 1,p (Ω) .
Then by the Gagliardo-Nirenberg inequality,

kuk np ≤ kEuk np ≤ CkD(Eu)kLp (Rn )


L n−p (Ω) L n−p (Rn )
≤ CkEukW 1,p (Rn ) ≤ CkukW 1,p (Ω) .

By following the proof of Theorem 2.89, we have the following general-


ization for integers k ≥ 1:

Theorem 2.97 (Gagliardo-Nirenberg inequality for W k,p (Ω)). Suppose that


Ω ⊂ Rn is open and bounded with C 1 boundary, 1 ≤ kp < n, and u ∈
W k,p (Ω). Then

kuk np ≤ CkukW k,p (Ω) for a constant C = C(k, p, n, Ω) .


L n−pk (Ω)

In fact, the theorem is true for real numbers s > 0 replacing integers k ≥
1, and follows from linear interpolation and the theory of fractional-order
Sobolev spaces defined later in Section 2.5.2. In the important case that
p = 2, we are then able to answer the question of which H s spaces embed
in Lq spaces. For example, when n = 2 and s = 12 , we see that kukL4 (Ω) ≤
Ckuk 1 , and when n = 3 and s = 21 , kuk 12 ≤ Ckuk 1 .
H 2 (Ω) L 5 (Ω) H 2 (Ω)

Theorem 2.98 (Gagliardo-Nirenberg inequality for W01,p (Ω)). Suppose that


Ω ⊂ Rn is open and bounded with C 1 boundary, 1 ≤ p < n, and u ∈ W01,p (Ω).
np
Then for all 1 ≤ q ≤ n−p ,

kukLq (Ω) ≤ CkDukLp (Ω) for a constant C = C(p, n, Ω) . (2.20)

Proof. By definition there exists a sequence uj ∈ C0∞ (Ω) such that uj → u


in W 1,p (Ω). Extend each uj by 0 on Ωc . Applying Theorem 2.89 to this
extension, and using the continuity of the norms, we obtain kuk pn ≤
L n−p (Ω)
CkDukLp (Ω) . Since Ω is bounded, the assertion follows by Hölder’s inequal-
ity. 
58 CHAPTER 2. INTRODUCTION TO SOBOLEV SPACES

Theorem 2.99. Suppose that Ω ⊂ R2 is open and bounded with C 1 bound-


ary, and u ∈ H01 (Ω). Then for all 1 ≤ q < ∞,

kukLq (Ω) ≤ C qkDukL2 (Ω) for a constant C = C(Ω) . (2.21)

Proof. The proof follows that of Theorem 2.90. Instead of introducing the
cut-off function g, we employ a partition of unity subordinate to the finite
covering of the bounded domain Ω, in which case it suffices that assume that
spt(u) ⊂ spt(U ) with U also defined in the proof Theorem 2.90. 

Remark 2.100. Inequalities (2.20) and (2.21) are commonly referred to


as Poincaré inequalities, and will be discussed in further detail in Theorem
5.22. They are invaluable in the study of the Dirichlet problem for Poisson’s
equation, since the right-hand side provides an H 1 (Ω)-equivalent norm for
all u ∈ H01 (Ω). In particular, there exists constants C1 , C2 such that

C1 kDukL2 (Ω) ≤ kukH 1 (Ω) ≤ C2 kDukL2 (Ω) .

2.2.11 Weak solutions to Dirichlet’s problem

Suppose that Ω ⊂ Rn is an open, bounded domain with C 1 boundary. A


classical problem in the linear theory of partial differential equations consists
of finding solutions to the Dirichlet problem:

−∆u = f in Ω , (2.22a)
u = 0 on ∂Ω , (2.22b)
Pn ∂2
where ∆ = i=1 ∂x2i denotes the Laplace operator or Laplacian. As written,
(2.22) is the so-called strong form of the Dirichlet problem, as it requires
that u to possess certain weak second-order partial derivatives. A major
turning-point in the modern theory of linear partial differential equations
was the realization that weak solutions of (2.22) could be defined, which only
require weak first-order derivatives of u to exist. (We will see more of this
idea later when we discuss the theory of distributions.)

Definition 2.101. The dual space of H01 (Ω) is denoted by H −1 (Ω). For
f ∈ H −1 (Ω),
kf kH −1 (Ω) = sup hf, ψi ,
kψkH 1 (Ω) =1
0
2.2. THE SOBOLEV SPACES H K (Ω) FOR INTEGERS K ≥ 0 59

where hf, ψi denotes the duality pairing between H −1 (Ω) and H01 (Ω).

Definition 2.102. A function u ∈ H01 (Ω) is a weak solution of (2.22) if


Z
Du · Dv dx = hf, vi ∀v ∈ H01 (Ω) .

Remark 2.103. Note that f can be taken in H −1 (Ω). According to the


Sobolev embedding theorem, this implies that when n = 1, the forcing func-
tion f can be taken to be the Dirac Delta distribution.

Remark 2.104. The motivation for Definition 2.102 is as follows. Since


C0∞ (Ω) is dense in H01 (Ω), multiply equation (2.22a) by φ ∈ C0∞ (Ω), inte-
R
grate over Ω, and employ the integration-by-parts formula to obtain Ω Du ·
R
Dφ dx = Ω f φ dx; the boundary terms vanish because φ is compactly sup-
ported.

Theorem 2.105 (Existence and uniqueness of weak solutions). For any


f ∈ H −1 (Ω), there exists a unique weak solution to (2.22).

Proof. Using the Poincaré inequality, kDukL2 (Ω) is an H 1 -equivalent norm


for all u ∈ H01 (Ω), and (Du, Dv)L2 (Ω) defines the inner-product on H01 (Ω).
As such, according to the definition of weak solutions to (2.22), we are
seeking u ∈ H01 (Ω) such that

(u, v)H01 (Ω) = hf, vi ∀v ∈ H01 (Ω) . (2.23)

The existence of a unique u ∈ H01 (Ω) satisfying (2.23) is provided by the


Riesz representation theorem for Hilbert spaces. 

Remark 2.106. Note that the Riesz representation theorem shows that there
exists a distribution, denoted by −∆u ∈ H −1 (Ω) such that

h−∆u, vi = hf, vi ∀v ∈ H01 (Ω) .

The operator −∆ : H01 (Ω) → H −1 (Ω) is thus an isomorphism.

A fundamental question in the theory of linear partial differential equa-


tions is commonly referred to as elliptic regularity, and can be explained as
follows: in order to develop an existence and uniqueness theorem for the
60 CHAPTER 2. INTRODUCTION TO SOBOLEV SPACES

Dirichlet problem, we have significantly generalized the notion of solution


to the class of weak solutions, which permitted very weak forcing functions
in H −1 (Ω). Now suppose that the forcing function is smooth; is the weak
solution smooth as well? Furthermore, does the weak solution agree with
the classical solution? The answer is yes, and we will develop this regular-
ity theory in Chapter 2.7, where it will be shown that for integers k ≥ 2,
−∆ : H k (Ω) ∩ H01 (Ω) → H k−2 (Ω) is also an isomorphism. An important
consequence of this result is that (−∆)−1 : H k−2 (Ω) → H k (Ω) ∩ H01 (Ω) is
a compact linear operator, and as such has a countable set of eigenvalues, a
fact that is eminently useful in the construction of solutions for heat- and
wave-type equations.
For this reason, as well as the consideration of weak limits of nonlinear
combinations of sequences, we must develop a compactness theorem, which
generalizes the well-known Arzela-Ascoli theorem to Sobolev spaces.

2.2.12 Strong compactness

In Section 2.1.12, we defined the notion of weak converence and weak com-
pactness for Lp -spaces. Recall that for 1 ≤ p < ∞, a sequence uj ∈ Lp (Ω)
converges weakly to u ∈ Lp (Ω), denoted uj * u in Lp (Ω), if Ω uj vdx →
R
R q p
Ω uvdx for all v ∈ L (Ω), with q = p−1 . We can extend this definition to
Sobolev spaces.

Definition 2.107. For 1 ≤ p < ∞, uj * u in W 1,p (Ω) provided that


uj * u in Lp (Ω) and Duj * Du in Lp (Ω).

Alaoglu’s Lemma (Theorem 2.37) then implies the following theorem.

Theorem 2.108 (Weak compactness in W 1,p (Ω)). For Ω ⊂ Rn , suppose


that
sup kuj kW 1,p (Ω) ≤ M < ∞ for a constant M 6= M (j) .

Then there exists a subsequence ujk * u in W 1,p (Ω).

It turns out that weak compactness often does not suffice for limit pro-
cesses involving nonlinearities, and that the Gagliardo-Nirenberg inequality
can be used to obtain the following strong compactness theorem.
2.2. THE SOBOLEV SPACES H K (Ω) FOR INTEGERS K ≥ 0 61

Theorem 2.109 (Rellich’s theorem on a bounded domain Ω). Suppose that


Ω ⊂ Rn is an open, bounded domain with C 1 boundary, and that 1 ≤ p < n.
np
Then W 1,p (Ω) is compactly embedded in Lq (Ω) for all 1 ≤ q < n−p , i.e. if

sup kuj kW 1,p (Ω) ≤ M < ∞ for a constant M 6= M (j) ,

then there exists a subsequence ujk → u in Lq (Ω). In the case that n = 2


and p = 2, H 1 (Ω) is compactly embedded in Lq (Ω) for 1 ≤ q < ∞.

In order to prove Rellich’s theorem, we need two lemmas.

Lemma 2.110 (Arzela-Ascoli Theorem). Suppose that uj ∈ C 0 (Ω), kuj kC 0 (Ω) ≤


M < ∞, and uj is equicontinuous. Then there exists a subsequence ujk → u
uniformly on Ω.

Lemma 2.111. Let 1 ≤ r ≤ s ≤ t ≤ ∞, and suppose that u ∈ Lr (Ω)∩Lt (Ω).


1 a 1−a
Then for s = r + t

kukLs (Ω) ≤ kukaLr (Ω) kuk1−a


Lt (Ω) .

Proof. By Hölder’s inequality,


Z Z
s
|u| dx = |u|as |u|(1−a)s dx
Ω Ω
Z  as Z  (1−a)s
r r t t
as as (1−a)s (1−a)s (1−a)s
≤ |u| dx |u| dx = kukas
Lr (Ω) kukLt (Ω) .
Ω Ω

Proof of Rellich’s theorem. Let Ω̃ ⊂ Rn denote an open, bounded domain


such that Ω ⊂⊂ Ω̃. By the Sobolev extension theorem, the sequence uj
satisfies spt(uj ) ⊂ Ω̃, and

sup kEuj kW 1,p (Rn ) ≤ CM .

Denote the sequence Euj by ūj . By the Gagliardo-Nirenberg inequality, if


np
1≤q< n−p ,

sup kukLq (Ω) ≤ sup kūkLq (Rn ) ≤ C sup kūj kW 1,p (Rn ) ≤ CM .
62 CHAPTER 2. INTRODUCTION TO SOBOLEV SPACES

For  > 0, let η denote the standard mollifiers and set ūj = η ∗ Euj .
By choosing  > 0 sufficiently small, ūj ∈ C0∞ (Ω̃). Since

1 y
Z Z
ūj = η( )ūj (x − y)dy = η(z)ūj (x − z)dz ,
B(0,) n  B(0,1)

and if ūj is smooth,


1 1
d
Z Z
ūj (x − z) − ūj (x) = ūj (x − tz)dt = − Dūj (x − tz) · z dt .
0 dt 0

Hence,
Z Z 1
|ūj (x) − ūj (x)| =  η(z) |Dūj (x − tz)| dzdt ,
B(0,1) 0

so that
Z Z Z 1Z
|ūj (x) − ūj (x)|dx =  η(z) |Dūj (x − tz)| dxdzdt
Ω̃ B(0,1) 0 Ω̃
≤ kDūj kL1 (Ω̃) ≤ kDūj kLp (Ω̃) < CM .

Using the Lp -interpolation Lemma 2.111,

kūj − ūj kLq (Ω̃) ≤ kūj − ūj kaL1 (Ω̃) kūj − ūj k1−a
np
L n−p (Ω̃)
1−a
≤ CM kDūj − Dūj kLp (Ω̃)

≤ CM M 1−a (2.24)

The inequality (2.24) shows that ūj is arbitrarily close to ūj in Lq (Ω)
uniformly in j ∈ N; as such, we attempt to use the smooth sequence ūj to
construct a convergent subsequence ūjk . Our goal is to employ the Arzela-
Ascoli Theorem, so we show that for  > 0 fixed,

kūj kC 0 (Ω̃) ≤ M̃ < ∞ and ūj is equicontinous.

For x ∈ Rn ,
Z
sup kūj kC 0 (Ω̃) ≤ sup η (x − y)|ūj (y)|dy
j j B(x,)
≤ kη kL∞ (Rn ) sup kūj kL1 (Ω̃) ≤ C−n < ∞ ,
j
2.2. THE SOBOLEV SPACES H K (Ω) FOR INTEGERS K ≥ 0 63

and similarly

sup kD̄uj kC 0 (Ω̃) ≤ kDη kL∞ (Rn ) sup kūj kL1 (Ω̃) ≤ C−n−1 < ∞ .
j j

The latter inequality proves equicontinuity of the sequence ūj , and hence
there exists a subsequence ujk which converges uniformly on Ω̃, so that

lim sup kūjk − ūjl kLq (Ω̃) = 0 .


k,l→∞

It follows from (2.24) and the triangle inequality that

lim sup kūjk − ūjl kLq (Ω̃) ≤ C .


k,l→∞

Letting C = 1, 21 , 13 , etc., and using the diagonal argument to extract


further subsequences, we can arrange to find a subsequence again denoted
by {ūjk } of {ūj } such that

lim sup kūjk − ūjl kLq (Ω̃) = 0 ,


k,l→∞

and hence
lim sup kujk − ujl kLq (Ω) = 0 ,
k,l→∞

The case that n = p = 2 follows from Theorem 2.90. 

Theorem 2.112. Let Ω denote an open, bounded, and smooth domain of


Rn , and let u ∈ H 1 (Ω) . Then u is absolutely continuous on almost all
straight lines parallel to the coordinate axes. Moreover, the weak derivatives
of u coincides with the classical derivative of u almost everywhere.

Proof. It suffices to assume that Ω = x ∈ Rn 0 < xi < 1 , 1 ≤ i ≤ n , and


show that
xn
∂u 0
Z
u(x) = (x , t)dt +const ,
∂xn
|0 {z }
≡v(x)

∂u
where the integrand is the weak derivative of u with respect to xn .
∂xn
64 CHAPTER 2. INTRODUCTION TO SOBOLEV SPACES

Let ω = x ∈ Rn−1 0 < xi < 1 , 1 ≤ i ≤ n − 1 so that Ω = ω × (0, 1) ,


and let ζ ∈ C0∞ (ω) and ϕ ∈ C0∞ (0, 1) be test functions. Since v is absolutely
continuous in xn , integration by parts implies
Z 1 Z 1
0 0
v(x , t)ϕ (t)dt = − vxn (x0 , t)ϕ(t)dt ,
0 0

where vxn denotes the classical derivative of v with respect to xn . Multi-


plying both sides by ζ(x0 ) and integrating over ω , we find that
Z Z
v(x)ζ(x )ϕ (xn )dx = − vxn (x)ζ(x0 )ϕ(xn )dx .
0 0
Ω Ω

By the definition of weak derivative,


∂u
Z Z
0 0
u(x)ζ(x )ϕ (xn )dx = − (x)ζ(x0 )ϕ(xn )dx .
Ω Ω ∂xn
∂u
Since the classical derivative vxn is the same as , the right-hand side of
∂xn
the two equalities above are the same; hence due to the fact that the test
function ζ ∈ C0∞ (ω) is arbitrary,
Z 1
u(x0 , xn ) − v(x0 , xn ) ϕ0 (xn )dxn = 0

0

for almost all x0 ∈ ω . As a consequence, by Problem 2.16, we find that

u(x0 , xn ) − v(x0 , xn ) = a constant independent of xn

which shows that u is absolutely continuous on almost all straight lines


parallel the xn -axis. 

2.2.13 Exercises

Problem 2.12. Suppose that 1 < p < ∞. If τy f (x) = f (x − y), show that
f belongs to W 1,p (Rn ) if and only if τy f is a Lipschitz function of y with
values in Lp (Rn ), i.e.

kτy f − τz f kLp (Rn ) ≤ C|y − z| .

What happens in the case p = 1?


2.2. THE SOBOLEV SPACES H K (Ω) FOR INTEGERS K ≥ 0 65

Problem 2.13. If for j = 1, 2 and pj ∈ [1, ∞] and uj ∈ Lpj , show that


u1 u2 ∈ Lr provided that 1/r = 1/p1 + 1/p2 and

ku1 u2 kLr ≤ ku1 kLp1 ku2 kLp2 .

Show that this implies that the generalized Hölder’s inequality, which states
that if for j = 1, ..., m and pj ∈ [1, ∞] with m 1
P
j=1 pj = 1, then
Z
|u1 · · · um | dx ≤ ku1 kLp1 · · · kum kLpm .
Rn

Problem 2.14. Let f ∈ L1 (R), and set


Z x
g(x) = f (y)dy . (*)
−∞

Continuity of g follows from the Dominated Convergence Theorem. Show


that ∂1 g = f .
(Hint. Given φ ∈ C0∞ (R), use (*) to obtain
Z Z Z x
φ0 (x)g(x)dx = φ0 (x)f (y)dydx .
R R −∞

Then write this integral as


x+h
1 1
Z Z Z
lim [φ(x + h) − φ(x)] g(x)dx = − lim f (y)φ(x) dydx .)
h→0 h R h→0 h R x

Problem 2.15. Show that W n,1 (Rn ) ⊂ C(Rn ) ∩ L∞ (Rn ).


R0 R0
(Hint. u(x) = −∞ · · · −∞ ∂1 · · · ∂n u(x + y)dy1 · · · dyn .)

Problem 2.16. If u ∈ W 1,p (Rn ) for some p ∈ [1, ∞) and ∂j u = 0 on a


connected open set Ω ⊂ Rn , for 1 ≤ j ≤ n, show that u is equal a.e. to a
constant on Ω.
(Hint. Approximate u using that

φi ∗ u → u in W 1,p (Rn ) ,
66 CHAPTER 2. INTRODUCTION TO SOBOLEV SPACES

where φi is a sequence of standard mollifiers. As we showed, given  > 0,


we can choose i such that

kφi ∗ u − ukW 1,p (Rn ) <  .

Show that ∂j (φi ∗ u) = 0 on Ωi ⊂⊂ Ω, where Ωi % Ω as i → ∞.)

More generally, if ∂j u = fj ∈ C(Ω), 1 ≤ j ≤ n, show that u is equal a.e.


to a function in C 1 (Ω).

Problem 2.17. In case n = 1, deduce from Problems 2.14 and 2.16 that, if
u ∈ L1loc (R) and if ∂1 u = f ∈ L1 (R), then
Z x
u(x) = c + f (y)dy , a.e. x ∈ R ,
−∞

for some constant c.

Problem 2.18. Let Ω := B(0, 12 ) ⊂ R2 denote the open ball of radius 1


2.
For x = (x1 , x2 ) ∈ Ω, let
q
u(x1 , x2 ) = x1 x2 log (| log(|x|)|) where |x| = x21 + x22 .

(a) Show that u ∈ C 1 (Ω̄);

∂2u
(b) show that ∂x2j
∈ C(Ω̄) for j = 1, 2, but that u 6∈ C 2 (Ω̄);

(c) show that u ∈ H 2 (Ω).

Problem 2.19. Theorem 2.84 states that for p > n, and y ∈ B(y, r) ,

1− n
|u(x) − u(y)| ≤ Cr p kDukLp (Rn ) ∀u ∈ C 1 (Rn ) . (2.25)

Prove that the inequality (2.25) in fact holds for all u ∈ W 1,p (Rn ); in par-
ticular, show that Du can be taken to be the weak derivative of u.
2.2. THE SOBOLEV SPACES H K (Ω) FOR INTEGERS K ≥ 0 67

Problem 2.20. Let η denote the standard mollifier, and for u ∈ H 2 (R3 ),
set u = η ∗ u. Prove that

ku − ukL∞ (R3 ) ≤ C kukH 2 (R3 ) ,

and that
ku − ukL∞ (R3 ) ≤ CkukH 3 (R3 ) .

Problem 2.21. Suppose that for n ≥ 2, Ω ⊂ Rn is a smooth, open, and


bounded domain, and let n denote the outward-pointing unit normal vector
to the boundary ∂Ω. Suppose that u ∈ L2 (Ω) and div u ∈ L2 (Ω). Prove that
u · n ∈ H −1/2 (∂Ω) and that

ku · nkH −1/2 (∂Ω) ≤ C kukL2 (Ω) + k div ukL2 (Ω) .

Problem 2.22. Let Ω ⊂ R2 denote an open, bounded, subset with smooth


boundary. Prove the interpolation inequality:

kDuk2L2 (Ω) ≤ CkukL2 (Ω) kD2 ukL2 (Ω) ∀ u ∈ H 2 (Ω) ∩ H01 (Ω) ,

where D2 u denotes the Hessian matrix of u, i.e., the matrix of second partial
∂2u
derivatives ∂xi ∂xj . Use the fact that C ∞ (Ω)∩H01 (Ω) is dense in u ∈ H 2 (Ω)∩
H01 (Ω).

Problem 2.23. Let D := B(0, 1) ⊂ R2 denote the unit disc, and let
h iα
u(x) = − log |x| .

Prove that the weak derivative of u exists for all α ≥ 0.

Problem 2.24. Suppose that {fn }∞ 1


n=1 is a bounded sequence in H (Ω) for
Ω ⊂ R2 bounded. Show that there exists an f ∈ H 1 (Ω) such that for 1 <
p < 2,
fnl Dfnl * f Df weakly in Lp (Ω) .

Problem 2.25. Suppose that uj * u in W 1,1 (0, 1). Show that uj → u a.e.
68 CHAPTER 2. INTRODUCTION TO SOBOLEV SPACES

2.3 The Fourier Transform

The Fourier transform is one of the most powerful and fundamental tools in
linear analysis, converting constant-coefficient linear differential operators
into multiplication by polynomials. In this section, we define the Fourier
transform, first on L1 (Rn ) functions, next (and miraculously) on L2 (Rn )
functions, and finally on the space of tempered distributions.

2.3.1 Fourier transform on L1 (Rn ) and the space S(Rn )

Definition 2.113. For all f ∈ L1 (Rn ) the Fourier transform F is defined


by Z
n
Ff (ξ) = fˆ(ξ) = (2π)− 2 f (x)e−ix·ξ dx .
Rn

By Hölder’s inequality, F : L1 (Rn ) → L∞ (Rn ).

Definition 2.114. The space of Schwartz functions of rapid decay is de-


noted by

S(Rn ) = {u ∈ C ∞ (Rn ) | xβ Dα u ∈ L∞ (Rn ) ∀α, β ∈ Zn+ }.

It is not difficult to show (as it follows from the definition) that

F : S(Rn ) → S(Rn ) ,

and that
ξ α Dξβ fˆ = (−i)|α| (−1)|β| F(Dxα xβ f ) .

The Schwartz space S(Rn ) is also known as the space of rapidly decreasing
functions; thus, after multiplying by any polynomial functions P (x), P (x)Dα u(x) →
0 as x → ∞ for all α ∈ Zn+ . The classical space of test functions D(Rn ) :=
2
C0∞ (Rn ) ⊂ S(Rn ). The prototype element of S(Rn ) is e−|x| which is not
compactly supported, but has rapidly decreasing derivatives.
The reader is encouraged to verify the following basic properties of S(Rn )
which we will denote by S:

1. S is a vector space.

2. S is an algebra under the pointwise product of functions.


2.3. THE FOURIER TRANSFORM 69

3. P (x)u(x) ∈ S for all u ∈ S and all polynomial functions P (x).

4. S is closed under differentiation.

5. S is closed under translations and multiplication by complex exponen-


tials eix·ξ .

6. S ⊂ L1 (Rn ) (since |u(x)| ≤ C(1 + |x|)n+1 for all u ∈ S and (1 +


|x|)−(n+1) dx decays like |x|−2 as |x| → ∞).

Definition 2.115. For all f ∈ L1 (Rn ), we define operator F ∗ by


Z
n
F ∗ f (x) = (2π)− 2 f (ξ)eix·ξ dξ .
Rn

Lemma 2.116. For all u, v ∈ S(Rn ),

(Fu, v)L2 (Rn ) = (u, F ∗ v)L2 (Rn ) .

Recall that the L2 (Rn ) inner-product for complex-valued functions is


R
given by (u, v)L2 (Rn ) = Rn u(x)v(x)dx.

Proof. Since u, v ∈ S(Rn ), by Fubini’s Theorem,


Z Z
−n
(Fu, v)L2 (Rn ) = (2π) 2 u(x)e−ix·ξ dx v(ξ) dξ
n n
ZR ZR
−n
= (2π) 2 u(x)eix·ξ v(ξ) dξ dx
n n
ZR R Z
−n
= (2π) 2 u(x) eix·ξ v(ξ) dξ dx = (u, F ∗ v)L2 (Rn ) ,
Rn Rn

Theorem 2.117. F ∗ ◦ F = Id = F ◦ F ∗ on S(Rn ).

Proof. We first prove that for all f ∈ S(Rn ), F ∗ Ff (x) = f (x).

Z Z 
∗ −n iξ·x −iy·ξ
F Ff (x) = (2π) e e f (y)dy dξ
n Rn
ZR Z
= (2π)−n ei(x−y)·ξ f (y) dy dξ .
Rn Rn

By the dominated convergence theorem,


Z Z
2
F ∗ Ff (x) = lim (2π)−n e−|ξ| ei(x−y)·ξ f (y) dy dξ .
→0 Rn Rn
70 CHAPTER 2. INTRODUCTION TO SOBOLEV SPACES

2
For all  > 0, the convergence factor e−|ξ| allows us to interchange the
order of integration, so that by Fubini’s theorem,
Z Z 
∗ −n −|ξ|2 i(y−x)·ξ
F Ff (x) = lim (2π) f (y) e e dξ dy .
→0 Rn Rn

Define the integral kernel


Z
−n 2 +ix·ξ
p (x) = (2π) e−|ξ| dξ
Rn

Then
Z

F Ff (x) = lim p ∗ f := p (x − y)f (y)dy .
→0 Rn
2 +ix·ξ
Let p(x) = p1 (x) = (2π)−n e−|ξ|
R
Rn dξ. Then

√ 2 +ix·ξ/√
Z
p(x/ ) = (2π)−n e−|ξ| dξ
n
ZR
2 +ix·ξ n n
= (2π)−n e−|ξ|  2 dξ =  2 p (x) .
Rn

We claim that
1
Z
|x|2
− 4
p (x) = n e and that p(x)dx = 1 . (2.26)
(4π) 2 Rn

0.9

0.8

0.7

0.6

0.5

0.4

0.3

0.2

0.1

0
-4 -2 0 2 4

Figure 2.1: As  → 0, the sequence of functions p becomes more localized


about the origin.

Given (2.26), then for all f ∈ S(Rn ), p ∗ f → f uniformly as  → 0,


which shows that F ∗ F = Id, and similar argument shows that FF ∗ = Id.
(Note that this follows from the proof of Theorem 2.28, since the standard
mollifiers η can be replaced by the sequence p and all assertions of the
2.3. THE FOURIER TRANSFORM 71

theorem continue to hold, for if (2.26) is true, then even though p does not
R
have compact support, B(0,δ)c p (x)dx → 0 as  → 0 for all δ > 0.)
Thus, it remains to prove (2.26). It suffices to consider the case  = 12 ;
then by definition
Z
|ξ|2
p 1 (x) = (2π)−n eix·ξ e− 2 dξ
2
Rn
 
|ξ|2
−n/2 − 2
= F (2π) e .

|x|2
In order to prove that p 1 (x) = (2π)−n/2 e− 2 , we must show that with the
2
|x|2
− 2
Gaussian function G(x) = (2π)−n/2 e ,

G(x) = F(G(ξ)) .

By the multiplicative property of the exponential,


2 /2 2 2
e−|ξ| = e−ξ1 /2 · · · e−ξn /2 ,

it suffices to consider the case that n = 1. Then the Gaussian satisfies the
differential equation
d
G(x) + xG(x) = 0 .
dx
Computing the Fourier transform, we see that
d
−i Ĝ(x) − iξ Ĝ(x) = 0 .

Thus,
ξ2
Ĝ(ξ) = Ce− 2 .
To compute the constant C,
Z
x2 1
−1
C = Ĝ(0) = (2π) e 2 dx = (2π)− 2
R
which follows from the fact that
Z
x2 1
e 2 dx = (2π) 2 . (2.27)
R
To prove (2.27), one can again rely on the multiplication property of the
exponential to observe that
Z x2 Z x2 Z x2 2
1 2 1 +x2
e 2 dx e 2 dx = e 2 dx
R R R2
Z 2π Z ∞
2
= e−2r rdrdθ = 2π .
0 0

72 CHAPTER 2. INTRODUCTION TO SOBOLEV SPACES

It follows from Lemma 2.116 that for all u, v ∈ S(Rn ),

(Fu, Fv)L2 (Rn ) = (u, F ∗ Fv)L2 (Rn ) = (u, v)L2 (Rn ) .

Thus, we have established the Plancheral theorem on S(Rn ).

Theorem 2.118 (Plancheral’s theorem). F : S(Rn ) → S(Rn ) is an iso-


morphism with inverse F ∗ preserving the L2 (Rn ) inner-product.

2.3.2 The topology on S(Rn ) and tempered distributions

An alternative to Definition 2.114 can be stated as follows:


p
Definition 2.119 (The space S(Rn )). Setting hxi = 1 + |x|2 ,

S(Rn ) = {u ∈ C ∞ (Rn ) | hxik |Dα u| ≤ Ck,α ∀k ∈ Z+ } .

The space S(Rn ) has a Fréchet topology determined by seminorms.

Definition 2.120 (Topology on S(Rn )). For k ∈ Z+ , define the semi-norm

pk (u) = sup hxik |Dα u(x)| ,


x∈Rn ,|α|≤k

and the metric on S(Rn )



X pk (u − v)
d(u, v) = 2−k .
1 + pk (u − v)
k=0

The space (S(Rn ), d) is a Fréchet space.

Definition 2.121 (Convergence in S(Rn )). A sequence uj → u in S(Rn ) if


pk (uj − u) → 0 as j → ∞ for all k ∈ Z+ .

Definition 2.122 (Tempered Distributions). A linear map T : S(Rn ) → C


is continuous if there exists some k ∈ Z+ and constant C such that

|hT, ui| ≤ Cpk (u) ∀u ∈ S(Rn ) .

The space of continuous linear functionals on S(Rn ) is denoted by S 0 (Rn ).


Elements of S 0 (Rn ) are called tempered distributions.

Definition 2.123 (Convergence in S 0 (Rn )). A sequence Tj * T in S 0 (Rn )


if hTj , ui → hT, ui for all u ∈ S(Rn ).
2.3. THE FOURIER TRANSFORM 73

For 1 ≤ p ≤ ∞, there is a natural injection of Lp (Rn ) into S 0 (Rn ) given


by Z
hf, ui = f (x)u(x)dx ∀u ∈ S(Rn ) .
Rn
Any finite measure on Rn provides an element of S 0 (Rn ). The basic example
of such a finite measure is the Dirac delta ‘function’ defined as follows:

hδ, ui = u(0) or, more generally, hδx , ui = u(x) ∀u ∈ S(Rn ) .

Definition 2.124. The distributional derivative D : S 0 (Rn ) → S(Rn ) is


defined by the relation

hDT, ui = −hT, Dui ∀u ∈ S(Rn ) .

More generally, the αth distributional derivative exists in S 0 (Rn ) and is de-
fined by
hDα T, ui = (−1)|α| hT, Dα ui ∀u ∈ S(Rn ) .

Multiplication by f ∈ S(Rn ) preserves S 0 (Rn ); in particular, if T ∈


S 0 (Rn ), then f T ∈ S 0 (Rn ) and is defined by

hf T, ui = hT, f ui ∀u ∈ S(Rn ) .

Example 2.125. Let H := 1[0,∞) denote the Heavyside function. Then


dH
= δ in S 0 (Rn ) .
dx
This follows since for all u ∈ S(Rn ),

dH du du
Z
h , ui = −hH, i = − dx = u(0) = hδ, ui .
dx dx 0 dx
Example 2.126 (Distributional derivative of Dirac measure).
dδ du
h , ui = − (0) ∀u ∈ S(Rn ) .
dx dx

2.3.3 Fourier transform on S 0 (Rn )

Definition 2.127. Define F : S 0 (Rn ) → S 0 (Rn ) by

hFT, ui = hT, Fui ∀u ∈ S(Rn ) ,

with the analogous definition for F ∗ : S 0 (Rn ) → S 0 (Rn ).


74 CHAPTER 2. INTRODUCTION TO SOBOLEV SPACES

Theorem 2.128. FF ∗ = Id = F ∗ F on S 0 (Rn ) .

Proof. By Definition 2.127, for all u ∈ S(Rn )

hFF ∗ T, ui = hF ∗ w, Fui = hT, F ∗ Fui = hT, ui ,

the last equality following from Theorem 2.117. 


n
Example 2.129 (Fourier transform of δ). We claim that Fδ = (2π)− 2 .
According to Definition 2.127, for all u ∈ S(Rn ),
Z
n
hFδ, ui = hδ, Fui = Fu(0) = (2π)− 2 u(x)dx ,
Rn
n
so that Fδ = (2π)− 2 .
n
Example 2.130. The same argument shows that F ∗ δ = (2π)− 2 so that
n n
F ∗ [(2π) 2 ] = 1. Using Theorem 2.128, we see that F(1) = (2π)− 2 δ. This
demonstrates nicely the identity

|ξ α û(ξ)| = |F(Dα u)(ξ)|.

In other the words, the smoother the function x 7→ u(x) is, the faster ξ 7→
û(ξ) must decay.

2.3.4 The Fourier transform on L2 (Rn )

In Theorem 2.28, we proved that C0∞ (Rn ) is dense in Lp (Rn ) for 1 ≤ p < ∞.
Since C0∞ (Rn ) ⊂ S(Rn ), it follows that S(Rn ) is dense in Lp (Rn ) as well.
Thus, for every u ∈ L2 (Rn ), there exists a sequence uj ∈ S(Rn ) such that
uj → u in L2 (Rn ), so that by Plancheral’s Theorem 2.118,

kûj − ûk kL2 (Rn ) = kuj − uk kL2 (Rn ) <  .

It follows from the completeness of L2 (Rn ) that the sequence ûj converges
in L2 (Rn ).

Definition 2.131 (Fourier transform on L2 (Rn )). For u ∈ L2 (Rn ) let uj


denote an approximating sequence in S(Rn ). Define the Fourier transform
as follows:
Fu = û = lim ûj .
j→∞
2.3. THE FOURIER TRANSFORM 75

Note well that F on L2 (Rn ) is well-defined, as the limit is independent


of the approximating sequence. In particular,

kûkL2 (Rn ) = lim kûj kL2 (Rn ) = lim kuj kL2 (Rn ) = kukL2 (Rn ) .
j→∞ j→∞

By the polarization identity

1 
(u, v)L2 (Rn ) = ku + vk2L2 (Rn ) − iku + ivk2L2 (Rn ) − (1 − i)kuk2L2 (Rn ) − (1 − i)kvk2L2 (Rn )
2

we have proved the Plancheral theorem1 on L2 (Rn ):

Theorem 2.132. (u, v)L2 (Rn ) = (Fu, Fv)L2 (Rn ) ∀u, v ∈ L2 (Rn ) .

2.3.5 Bounds for the Fourier transform on Lp (Rn )


n
We have shown that for u ∈ L1 (Rn ), kûkL∞ (Rn ) ≤ (2π)− 2 kukL1 (Rn ) , and
that for u ∈ L2 (Rn ), kûkL2 (Rn ) = kukL2 (Rn ) . Interpolating p between 1 and
2 yields the following result.

Theorem 2.133 (Hausdorff-Young inequality). If u ∈ Lp (Rn ) for 1 ≤ p ≤


p−1
2, then for q = p , there exists a constant C such that

kûkLq (Rn ) ≤ CkukLp (Rn ) .

Returning to the case that u ∈ L1 (Rn ), not only is Fu ∈ L∞ (Rn ), but


the transformed function decays at infinity.

Theorem 2.134 (Riemann-Lebesgue “lemma”). For u ∈ L1 (Rn ), Fu is


continuous and Fu(ξ) → 0 as |ξ| → ∞.

Proof. Let BM = B(0, M ) ⊂ Rn . Since f ∈ L1 (Rn ), for each  > 0, we can


choose M sufficiently large such that fˆ(ξ) ≤  + BM e−ix·ξ |f (x)|dx. Using
R

Lemma 2.23, choose a sequence of simple functions φj (x) → f (x) a.e. on


BM . For jnN chosen sufficiently large,
Z
ˆ
f (ξ) ≤ 2 + φj (x)e−ix·ξ dx .
BM

1
The unitarity of the Fourier transform is often called Parseval’s theorem in science
and engineering fields, based on an earlier (but less general) result that was used to prove
the unitarity of the Fourier series.
76 CHAPTER 2. INTRODUCTION TO SOBOLEV SPACES

PN
Write φj (x) = l=1 Cl 1El (x) so that
N Z
fˆ(ξ) ≤ 2 +
X
Cl φj (x)e−ix·ξ dx .
l=1 El

By the regularity of the Lebesgue measure µ, for all  > 0 and each l ∈
{1, ..., N }, there exists a compact set Kl and an open set Ol such that

µ(Ol ) − /2 < µ(El ) < µ(Kl ) + /2 .

Then Ol = {∪α∈Al Vαl | Vlα ⊂ Rn is open rectangle , Al arbitrary set }, and


Kl ⊂ ∪N l
j=1 Vj ⊂ Ol where {1, ..., Nl } ⊂ Al such that
l

|µ(El ) − µ(∪N l
j=1 Vj )| <  .
l

It follows that Z Z

e−ix·ξ dx − N e−ix·ξ dx <  .


El l
∪j=1 Vjl
On the other hand, for each rectangle Vjl , V l e−ix·ξ dx| ≤ C/(ξ1 · · · ξn ), so
R
j
that  
1
fˆ(ξ) ≤ C  + .
ξ1 · · · ξn
Since  > 0 is arbitrary, we see that fˆ(ξ) → 0 as |ξ| → ∞. Continuity of Fu
follows easily from the dominated convergence theorem. 

2.3.6 Convolution and the Fourier transform

Theorem 2.135. If u, v ∈ L1 (Rn ), then u ∗ v ∈ L1 (Rn ) and


n
F(u ∗ v) = (2π) 2 Fu Fv .

Proof. Young’s inequality (Theorem 2.53) shows that u ∗ v ∈ L1 (Rn ) so


that the Fourier transform is well-defined. The assertion then follows from
a direct computation:
Z
n
F(u ∗ v) = (2π)− 2 e−ix·ξ (u ∗ v)(x)dx
Rn
Z Z
n
= (2π)− 2 u(x − y)v(y)dy e−ix·ξ dx
n n
ZR ZR
n
= (2π)− 2 u(x − y)e−i(x−y)·ξ dx v(x) e−iy·ξ dy
Rn Rn
n
= (2π) ûv̂ (by Fubini’s theorem) .
2


2.3. THE FOURIER TRANSFORM 77

By using Young’s inequality (Theorem 2.54) together with the Hausdorff-


Young inequality, we can generalize the convolution result to the following

Theorem 2.136. Suppose that u ∈ Lp (Rn ) and v ∈ Lq (Rn ), and let r


r
1 1 1
satisfy r = p + q − 1 for 1 ≤ p, q, r ≤ 2. Then F(u ∗ v) ∈ L r−1 (Rn ) and
n
F(u ∗ v) = (2π) 2 Fu Fv .

2.3.7 An explicit computation with the Fourier Transform

The computation of the Green’s function for the Laplace operator is an


important application of the Fourier transform. For this purpose, we will
compute fˆ for the following two cases: (1) f (x) = e−t|x| , t > 0 and (2)
f (x) = |x|α , −n < α < 0.
Case (1) In this case, f is rapidly decreasing but not in the Schwartz class
S(Rn ). We begin with n = 1. It follows that
Z ∞ Z 0 Z ∞
−t|x| −t|x| −ix·ξ x(t−iξ)
[F(e )](ξ) = e e dµ1 (x) = e dµ1 (x) + ex(−t−iξ) dµ1 (x)
−∞ −∞ 0
r
1 e
h x(t−iξ) 0 e x(−t−iξ) ∞ i 2 t
=√ + = .

2π t − iξ −∞ −t − iξ 0 π t + ξ2
2

1 ∞ t
Z
−t|x|
By the inversion formula, we then see that e = eixξ dξ.
π −∞ t + ξ 2
2
Next, when n > 1 we will show that for some function g(t, s),
Z ∞
−t|x| 2
e = g(t, s)e−s|x| ds . (2.28)
0

In order to determine g(t, s), we suppose that (2.28) holds, and compute its
Fourier transform:
Z ∞ Z ∞  1 n −|ξ|2
−t|x| −s|x|2
F(e )= g(t, s)F(e )ds = g(t, s) √ e 4s ds ,
0 0 2s
where we have used the definition of the Fourier transform of the Gaussian
function given in the proof of Theorem 2.117. We are thus seeking a function
g(t, s) which satisfies
Z ∞
−tλ 2
e = g(t, s)e−sλ ds, ∀ λ > 0.
0
We begin by computing
Z ∞ 2 2
−st2 −sξ 2 e−s(t +ξ ) ∞ 1
e e ds = 2 2 = 2 . (2.29)
0 −(t + ξ ) 0 t + ξ2
78 CHAPTER 2. INTRODUCTION TO SOBOLEV SPACES

With λ = |x| > 0, we use (2.29) to find that


1 ∞ t 1 ∞  ∞ −st2 −sξ2  iλξ
Z Z Z
−tλ iλξ
e = e dξ = t e e ds e dξ
π −∞ t2 + ξ 2 π −∞ 0
1 ∞  ∞ −sξ2 iλξ  −st2
Z ∞
t
Z Z
2 |x|2
= t e e dξ e ds = √ e−st e− 4s ds
π 0 −∞ 0 sπ
so that
√ n t 2
g(t, s) = 2s √ e−st ,

and hence
∞ ∞
t
Z Z
|x|2n 2 2
−t|x| −
F(e )(ξ) = g(t, s)F(e √ (2s) 2 e−s(t +|ξ| ) ds
4s )ds =
0 0 sπ

t 1 C(n)t
Z
n
= n+1 √ (2s) 2 e−s ds = n+1 ,
(t2 + |ξ|2 ) 2 0 πs (t2 + |ξ|2 ) 2
Z ∞ r
1 n 2n n + 1
where the constant C(n) = √ (2s) 2 ds = Γ( ), and Γ is the
0 πs π 2
so-called gamma-function. It follows that
r
2n n + 1 t
F −1 (e−t|ξ| )(x) = F(e−t|ξ| )(−x)= Γ( ) n+1 . (2.30)
π 2 (t + |x|2 ) 2
2

Case (2) For this case, we compute F(|x|α ), when −n < α < 0. Using the
definition of Γ(n) above, we see that
Z ∞ ∞
α
Z
−α −1 −s|x| 2 α
s 2 e ds = |x| α
s− 2 −1 e−s ds = |x|α Γ(− ) ,
0 0 2
Therefore,
∞ ∞
1 1
Z Z
|ξ|2
−α −1 −s|x|2 α n
α
F(|x| ) = s 2 F(e )ds = n s− 2 − 2 −1 e− 4s ds
Γ(− α2 ) 0 2 Γ(− α2 )
2 0
n
1 ∞
|ξ|2 − α2 − n2 −1 −s |ξ|2 2α+ 2 Γ( α+n
2 )
Z
= n e 2
ds = α |ξ|−α−n ,
2 2 Γ(− α2 ) 0 4s 4s Γ(− 2 )

where we impose the condition −n < α < 0 to ensure the boundedness of


the Γ-function. In particular, for n = 3 and α = −1,
√ r
−1 2Γ(1) −2 2 −2
F(|x| ) = 1 |ξ| = |ξ| ,
Γ( 2 ) π

from which it follows that


r
−1 −2 π 1
F (|ξ| )= . (2.31)
2 |x|
2.3. THE FOURIER TRANSFORM 79

2.3.8 Applications to the Poisson, Heat, and Wave equations


The Poisson equation on R3

In Theorem 2.105, we proved the existence of unique weak solutions to the


Dirichlet problem on a bounded domain Ω. We will now provide an explicit
representation for solutions to the Poisson problem on R3 . The issue of
uniqueness in this setting will be of interest.
Given the Poisson problem

∆u = f in S 0 ,

we compute the Fourier transform of both sides to obtain that

−|ξ|2 û(ξ) = fˆ(ξ) . (2.32)

Distributional solutions to (2.32) are not unique; for example,

fˆ(ξ) fˆ(ξ)
û(ξ) = − and û(ξ) = − +δ
|ξ|2 |ξ|2
are both solutions. By requiring solutions to have enough decay, such as
u ∈ L2 (Rn ) so that û ∈ L2 (Rn ), then we do obtain uniqueness.
We will find an explicit representation for the solution to the Poisson
problem when n = 3. If u ∈ L2 (R3 ), then using (2.31), we see that

fˆ(ξ)  fˆ(ξ) 
û(ξ) = − 2 ⇒ u(x) = −F −1 (x)
|ξ| |ξ|2
h i
= − F −1 (|ξ|−2 ) ? F −1 (fˆ) (x) = (Φ ∗ f )(x) ,

1
where Φ(x) = − . The function Φ is the so-called fundamental solution;
4π|x|
more precisely, it is the distributional solution of the equation

∆Φ = δ in S 0 .

Conceptually

∆(Φ ∗ f ) = ∆Φ ∗ f = δ∗f = f ∀ f ∈ C(Rn ) whenever Φ ∗ f makes sense ,

where the first equality follows from the fact that


n n
h∆(Φ ∗ f ), ϕ̂i = (2π) 2 h−|ξ|2 Φ̂fˆ, ϕi = (2π) 2 hF(∆Φ), fˆϕi
n
= (2π) 2 h∆Φ, F(fˆϕ)i = h∆Φ, fe ∗ ϕ̂i = h∆Φ ∗ f, ϕ̂i .
80 CHAPTER 2. INTRODUCTION TO SOBOLEV SPACES

Example 2.137. On R2 , ∆(ex1 cos x2 ) = 0. The function ex1 cos x2 is not


a tempered distribution because it grows too fast as x1 → ∞. As such, the
Fourier transfor of ex1 cos x2 is not defined.
Using Fourier transform to convert PDE to linear algebraic equations
only provides those solutions which do not grow too rapidly at ∞.

The Poisson integral formula on the half-space

Let Ω = Rn × R+ =, and consider the Dirichlet problem


h ∂2 ∂2 ∂2 i h ∂2 i
+ + · · · + u = + ∆ u = 0 in Ω × (0, ∞) ,
∂t2 ∂x21 ∂x2n ∂t2
u(x, 0) = f (x)(∈ S(Rn )) on ∂Ω × (0, ∞) .

(Note that for any constant c, c t is always a solution as it is harmonic


and vanishes at the boundary t = 0.) For uniqueness, we insist that u be
bounded. This in turn means u is in S 0 and hence we may use the Fourier
transform. Applying the Fourier transform (in the x variable) Fx , we see
that
∂2
Fx u(ξ, t) − |ξ|2 Fx u(ξ, t) = 0 , Fx u(ξ, 0) = fˆ(ξ) .
∂t2
Therefore, Fx u(ξ, t) = C1 (ξ)et|ξ| + C2 (ξ)e−t|ξ| , and C1 (ξ) = 0 by the growth
condition imposed on u. Then Fx u(ξ, t) = fˆ(ξ)e−t|ξ| and hence using (2.30),
h i
u(x, t) = F −1 (fˆ(ξ)e−t|ξ| )(x) = F −1 (e−t|ξ| ) ∗ f (x)
Γ( n+1
2 ) tf (y)
Z
= n+1 n+1 dy .
π 2 Rn (t2 + |x − y|2 ) 2

This is the Poisson integral formula on the half-space.


If f is bounded, i.e., f ∈ L∞ (Rn ), then the integral converges and u ∈
L∞ (Rn × R+ ). Therefore, u ∈ C ∞ (Rn × R+ ) ∩ L∞ (Rn × R+ ).

The Heat equation

Let t ≥ 0 denote time, and x denote a point in space Rn . The function u(x, t)
denotes the temperature at time t and position x, and g ∈ S(Rn ) denotes
the initial temperature distribution. We wish to solve the heat equation

ut (x, t) = ∆u(x, t) in Rn × (0, ∞) , (2.33a)


u(x, 0) = g(x) on Rn × {t = 0} . (2.33b)
2.3. THE FOURIER TRANSFORM 81

Taking the Fourier transform of (2.33), we find that

∂t û(ξ, t) = −|ξ|2 û(ξ, t) ,


û(ξ, 0) = ĝ(ξ) .

2t
Therefore, û(ξ, t) = ĝ(ξ)e−|ξ| and hence
 2
 h  2
 i
u(x, t) = F −1 ĝ(ξ)e−|ξ| t (x) = F −1 e−|ξ| t ∗ g (x)
1
Z
|x−y|2
− 4t
= e g(y)dy (≡ (H(·, t) ∗ g)(x)) . (2.34)
(4πt)n/2 Rn

Theorem 2.138. If g ∈ L∞ (Rn ), then the solution u to (2.33) is in C ∞ (Rn ×


(0, ∞)).
2
e−|x| /4t
Proof. The function is C ∞ (Rn × [α, ∞)) for all α > 0. 
(4πt)n/2

Remark 2.139. The representation formula (2.34) shows that whenever g


is bounded, continuous, and positive, the solution u(x, t) to (2.33) is positive
everywhere for t > 0.

The representation formula (2.34) can also be used to prove the following

Theorem 2.140. Assume that g ∈ C(Rn ) ∩ L∞ (Rn ). Then u defined by


(2.34) is continuous at t = 0, that is,

lim u(x, t) = g(x0 ) ∀ x0 ∈ Rn .


(x,t)→(x0 ,0+ )

In order to study the Inhomogeneous heat equation

ut (x, t) − ∆u(x, t) = f (x, t) in Rn × (0, ∞) , (2.35a)


u(x, 0) = 0 on Rn × {t = 0} , (2.35b)

we introduce the parameter s > 0, and consider the following problem for
U:

Ut (x, t, s) = ∆U (x, t, s) ,
U (x, s, s) = f (x, s) .
82 CHAPTER 2. INTRODUCTION TO SOBOLEV SPACES

Then by (2.34),
Z
U (x, t, s) = H(x − y, t − s)f (y, s)dy .
Rn

We next invoke Duhamel’s principle to find a solution u(x, t) to (2.35):


Z t Z tZ
u(x, t) = U (x, t, s)ds = H(x − y, t − s)f (y, s)dyds . (2.36)
0 0 Rn

The principle of linear superposition then shows that the solution of the
problem

ut (x, t) − ∆u(x, t) = f (x, t) in Rn × (0, ∞) ,


u(x, 0) = g(x) on Rn × {t = 0} ,

is the sum of (2.34) and (2.36):


Z tZ Z
u(x, t) = H(x − y, t − s)f (y, s)dyds + H(x − y, t)g(y)dy
0 Rn Rn
Z t
= [H(·, t) ∗ g](x) + [H(·, t − s) ∗ f (·, s)](x)ds . (2.37)
0

The Wave equation

For wave speed c > 0, and for x ∈ Rn , t ∈ R, consider the following second-
order linear hyperbolic equation:

utt (x, t) = c2 ∆u(x, t) in Rn × (0, ∞) ,


u(x, 0) = f (x) on Rn × {t = 0} ,
ut (x, 0) = g(x) on Rn × {t = 0} .

Taking the Fourier transform of (2.38), we find that

ûtt (ξ, t) = −c2 |ξ|2 û(ξ, t) in Rn × (0, ∞) ,


û(ξ, 0) = fˆ(ξ) on Rn × {t = 0} ,
ût (ξ, 0) = ĝ(ξ) on Rn × {t = 0} .

The general solution of this second-order ordinary differential equations is


given by
û(ξ, t) = C1 (ξ) cos c|ξ|t + C2 (ξ) sin c|ξ|t .
2.3. THE FOURIER TRANSFORM 83

Solving for C1 and C2 by using the initial conditions, we find that

sin c|ξ|t
û(ξ, t) = fˆ(ξ) cos c|ξ|t + ĝ(ξ) .
c|ξ|

Therefore,
h  sin c|ξ|t  i
u(x, t) = F −1 (cos c|ξ|t) ∗ f + F −1 ∗ g (x)
c|ξ|
1 h d −1  sin c|ξ|t   sin c|ξ|t  i
= F ∗ f + F −1 ∗ g (x) .
c dt |ξ| |ξ|

For the case that n = 1,

m
sin ctλ −ixλ m
ei(ct−x)λ − e−i(ct+x)λ
Z Z
e dλ = dλ .
−m λ −m 2iλ

m
eiz
Z
By the Cauchy integral formula and the residue theorem, lim dz = iπ .
m→∞ −m z
Therefore, ∀ t > 0,

m 
1 sin ctλ −ixλ
Z
1 |x| < ct
lim e dλ = χ|x|<ct (x) = .
m→∞ π −m λ 0 |x| ≥ ct

 sin c|ξ|t  r
−1 π
Corollary 2.141. F (x) = χ (x) in S 0 (R).
|ξ| 2 |x|<ct

Proof. For all ϕ ∈ S(R),


Z  sin c|ξ|t  Z
sin c|ξ|t −1
−1
F (x)ϕ(x)dx = F (ϕ)(ξ)dξ
R |ξ| R |ξ|
Z mZ
1 sin c|ξ|t ixξ
= lim √ e ϕ(x)dxdξ ,
m→∞ 2π −m R |ξ|

and by Fubini’s theorem together with the dominated convergence theorem,


we see that
m Z Z m
sin c|ξ|t ixξ sin c|ξ|t ixξ
Z Z
lim e ϕ(x)dxdξ = lim e ϕ(x)dξdx
m→∞ −m R |ξ| m→∞ R −m |ξ|
Zˇ Z m
sin c|ξ|t ixξ
Z
= lim e ϕ(x)dξdx = π χ|x|<ct (x)ϕ(x)dx .
R m→∞ −m |ξ| R


84 CHAPTER 2. INTRODUCTION TO SOBOLEV SPACES

We have thus established d’Alembert’s formula for the solution of the


the 1-D wave equation:
r
1 π 1 hd
Z Z i
u(x, t) = √ f (x − y)χ|y|<ct (y)dy + g(x − y)χ|y|<ct (y)dy
c 2 2π dt R R
Z ct Z ct
1 d 1
= f (x − y)dy + g(x − y)dy
2c dt −ct 2c −ct
f (x − ct) + f (x + ct) 1 x+ct
Z
= + g(y)dy .
2 2c x−ct

We have just used the Fourier transform to find explicit solutions to


the fundamental linear elliptic, parabolic, and hyperbolic equations. More
generally, the Fourier transform is a powerful tool for the analysis of many
other constant coefficient linear partial differential equations.

2.3.9 Exercises

Problem 2.26. The general solution u(x, t) to the 1-D wave equation

utt (x, t) = c2 uxx (x, t) in R × (0, ∞)

can be expressed as u(x, t) = F (x + ct) + G(x − ct) for some functions F


and G. Use this expression to solve the initial-value problem

utt (x, t) = c2 uxx (x, t) in R × (0, ∞) ,


u(x, 0) = f (x) on R × {t = 0} ,
ut (x, 0) = g(x) on R × {t = 0} .

Check if your answer agrees with the d’Alembert formula.

Problem 2.27. (a) Using the Euler identity, compute F −1 (cos c|ξ|t) di-
 sin c|ξ|t 
rectly for the case n = 1. Find F −1 via the formula
|ξ|

1 d −1  sin c|ξ|t 
F −1 (cos c|ξ|t) = F .
c dt |ξ|
2.3. THE FOURIER TRANSFORM 85

(b) For the case that n = 3, compute

−1 sin c|ξ|t sin c|ξ|t ix·ξ


  Z
F (x) = lim e dµ3 (ξ) .
|ξ| m→∞ |ξ|≤m |ξ|
Using spherical coordinates, show that
 1 3 Z m Z π Z 2π sin cρt
−1 sin c|ξ|t
 
F (x) = lim √ ei|x|ρ cos φ ρ2 sin φdθdφdρ
|ξ| m→∞ 2π ρ
Z m 0 0 0
1 sin cρt i|x|ρ cos φ φ=π
=√ lim e dρ
2π m→∞ 0 −i|x|

φ=0
Z m −icρt
1 e − eicρt h i|x|ρ i
=√ lim e − e−i|x|ρ dρ
2π m→∞ 0 2|x|
Z ∞h
1 i
= eiρ(|x|−ct) − eiρ(|x|+ct) dµ1 (ρ) .
2|x| −∞

(c) For t > 0 ,


h  sin c|ξ|t  i
F −1 ? ϕ (x)
|ξ|
Z ∞Z ∞
1 1 iρ(r−ct)
Z
= e ϕ(x − rω)r2 dµ1 (ρ)dµ1 (r)dSω
4π ∂B(0,1) 0 −∞ r
ct 1
Z Z
= ϕ(x − ctw)dSω = ϕ(y)dSy ,
4π ∂B(0,1) 4πct ∂B(x,ct)
and conclude that
1 ∂h 1
Z Z i
u(x, t) = g(y)dSy + f (y)dSy .
4πc2 t ∂B(x,ct) ∂t 4πc2 t ∂B(x,ct)
(2.39)

Problem 2.28. Derive the solution to the 2-D and 3-D wave equations

utt (x, t) = c2 ∆u(x, t) in Rn × (0, ∞) , (2.40a)


u(x, 0) = f (x) on Rn × {t = 0} , (2.40b)
ut (x, 0) = g(x) on Rn × {t = 0} , (2.40c)

by the method of spherical means.


Let Mh (x, r) denote the average of the function h on a sphere with center
x and radius r (hence x and r are independent variable here), that is,
1
Z
Mh (x, r) = h(y)dSy . (2.41)
ωn−1 rn−1 ∂B(x,r)
86 CHAPTER 2. INTRODUCTION TO SOBOLEV SPACES

Setting y = x + rω with |ω| = 1 , we get


1
Z
Mh (x, r) = h(x + rω)dSω . (2.42)
ωn−1 ∂B(0,1)

Note that in (2.41) Mh (x, r) is defined only for r > 0 , while we can extend
its definition to all r ∈ R using (2.42). Therefore, we treat Mh (x, r) as a
function on Rn+1 . If h ∈ C k (Rn ) , then Mh ∈ C k (Rn+1 ) . For h ∈ C 2 (Rn ) ,
n
∂ 1
Z X
Mh (x, r) = hxi (x + rω)ωi dSω
∂r ωn−1 ∂B(0,1) i=1
r
Z
by divergence theorem = ∆x h(x + ry)dy
ωn−1 B(0,1)
Z rZ
r1−n
= ∆x h(x + ρω)ρn−1 dSω dρ
ωn−1 0 ∂B(0,1)
Z r
1−n
=r ∆x ρn−1 Mh (x, ρ)dρ .
0

Therefore, for all h ∈ C 2 (Rn ) ,


∂ h n−1 ∂ i
r Mh (x, r) = rn−1 ∆x Mh (x, r)
∂r ∂r
h ∂2 n−1 ∂ i
⇒ + Mh (x, r) = ∆x Mh (x, r) .
∂r2 r ∂r
The last equality is called the Darboux equation. By (2.42), Mh (x, r) is even

in r , so Mh (x, 0) = h(x) and Mh (x, r) = 0.

∂r r=0

Now let u be a solution to (2.40), and let Mu (x, r, t) be the spherical


mean of u , that is,
1
Z
Mu (x, r, t) = u(x + rω, t)dSω .
ωn−1 ∂B(0,1)

Then Mu (x, 0, r) = u(x, t) . Moreover,


1
Z
∆x M u = ∆x u(x + rω, t)dSω
ωn−1 ∂B(0,1)
1 ∂2 h 1 1 ∂2
Z i
= 2 2 u(x + rω, t)dSω = 2 2 Mu .
c ∂t ωn−1 ∂B(0,1) c ∂t

By the Darboux equation,

∂2 h 2
2 ∂ n−1 ∂ i
M u = c + Mu . (2.43)
∂t2 ∂r2 r ∂r
2.3. THE FOURIER TRANSFORM 87

We then transform a n + 1-dimensional PDE to a 1 + 1-dimensional PDE.


n = 3: Let n = 3 in (2.43), we find that
∂2 2 ∂
2
(rM u ) = c (rMu ) .
∂t2 ∂r2
By the d’Alembert formula (here r plays the role of x in the d’Alembert
formula), we find that
1h i
Mu (x, r, t) = (r + ct)Mf (x, r + ct) + (r − ct)Mf (x, r − ct)
2r
Z r+ct
1
+ sMg (x, s)ds .
2cr r−ct
Use lim Mu (x, r, t) = u(x, t) to find u . Does this agree with (2.39)?
r→0
Problem 2.29. By (2.39), the solution of the 3-dimensional wave equation

utt (x, t) = c2 ∆u(x, t) in R3 × (0, ∞) , (2.44a)


u(x, 0) = f (x) on R3 × {t = 0} , (2.44b)
ut (x, 0) = g(x) on R3 × {t = 0} , (2.44c)

can be expressed as
1 ∂h 1
Z Z i
u(x, t) = g(y)dS y + f (y)dSy .
4πc2 t ∂B(x,ct) ∂t 4πc2 t ∂B(x,ct)
Suppose that f ∈ C02 (R3 ) and g ∈ C01 (R3 ) so that they provide a solution
u ∈ C 2 (R3 × (0, ∞)) . Show that there exists a constant K > 0 so that
K
|u(x, t)| ≤ ∀t > 0.
t
Hint: First rewrite (2.39) as
3
1
Z h X i
u(x, t) = 2 2 2 tg(y) + f (y) + fyi (y)(yi − xi ) dSy
4π c t |y−x|=ct
i=1
and convert the integral into an integral over B(x, ct) .

Proof. By the change of variable y = x + ctω ,


∂ ∂
Z Z
f (y)dSy = f (x + ctω)c2 t2 dSω
∂t ∂B(x,ct) ∂t ∂B(0,1)
Z h i
= 2c2 tf (x + ctω) + (Df )(x + ctω)c3 t2 ωi dSω
∂B(0,1)
1h
Z i
= 2f (y) + (Df )(y) · (y − x) dSy ,
∂B(x,ct) t
88 CHAPTER 2. INTRODUCTION TO SOBOLEV SPACES

we find that
3
1
Z h X i
u(x, t) = 2 2 2 tg(y) + f (y) + fyi (y)(yi − xi ) dSy
4π c t |y−x|=ct i=1
3 Z
1 X h tg(y) + f (y) i
= (yi − xi ) + ctfyi (y) Ni dSy
4π 2 c2 t2 ∂B(x,ct) ct
i=1
1
Z h 3tg(y) + 3f (y) + t(Dg)(y) + (Df )(y) · (y − x) i
= 2 2 2 + ct(∆f )(y) dy ;
4π c t B(x,ct) ct

hence
C |f (y)|
Z h i
|u(x, t)| ≤ |g(y)| + t|Dg(y)| + + |Df (y)| + t|∆f (y)| dy
t2 B(x,ct) t
Ch i
≤ kgk 23 3 + kDgkL1 (R3 ) + kf kL3 (R3 ) + kDf k 23 3 + k∆f kL1 (R3 ) .
t L (R ) L (R )

Problem 2.30. Let us consider the BBM equation

ut + ux + uux − uxxt = 0 ∀ x ∈ R , t ∈ (0, T ] , (2.45a)


u(x, 0) = g(x) ∀x ∈ R. (2.45b)

(1) Use the Fourier transform to show that a bounded solution to (2.45)
satisfies
Z tZ ∞ h 1 i
u(x, t) = g(x) + K(x − y) u(y, s) + u2 (y, s) dyds , (2.46)
0 −∞ 2

where K is defined by
1
K(x) = sgn(x)e−|x| .
2

(2) Write (2.46) as u = F (u) , that is, treat the right-hand side of (2.46)
as a function of u . Show that for T > 0 small enough, F has a fixed-
point in the space of bounded continuous functions. (Hint: similar to
the proof of the fundamental theorem of ODE, you can try to show that
the map F is a contraction mapping if T is small enough, and then
apply the contraction mapping theorem.)
2.3. THE FOURIER TRANSFORM 89

1 RR
Problem 2.31. (a) For f ∈ L1 (R), set SR f (x) = (2π)− 2 ˆ ixξ dξ.
−R f (ξ)e
Show that Z ∞
SR f (x) = KR ∗ f (x) = KR (x − y)f (y)dy
−∞

where
R
sin Rx
Z
KR (x) = (2π)−1 eixξ dξ = .
−R πx

(b) Show that if f ∈ L2 (R), then SR f → f in L2 (R) as R → ∞.

Problem 2.32. Show that for any R ∈ (0, ∞), there exists f ∈ L1 (R) such
that SR f 6∈ L1 (R). (Hint. Note that KR 6∈ L1 (R).) For partial credit,
explain why the result is interesting.

Problem 2.33. Assume w ∈ S 0 ∩ L1loc (Rn ) and w(x) ≥ 0. Show that if


ŵ ∈ L∞ (Rn ), then w ∈ L1 (Rn ) and

kŵkL∞ (Rn ) = (2π)−n/2 kwkL1 (Rn )

(Hint. Consider wj (x) = ψ( xj )w(x) with ψ ∈ C0∞ (Rn ) and ψ(0) = 1. Use
the fact that wj * w in S 0 .)

Problem 2.34. Consider the Poisson equation on R1 : uxx = f .


x + |x| |x|
(a) Show that ϕ(x) = and ψ(x) = are both distributional solu-
2 2
tions to uxx = δ0 .

(b) Let f be continuous with compact support in R . Show that


Z
u(x) = ϕ(x − y)f (y)dy
R

and Z
v(x) = ψ(x − y)f (y)dy
R

both solve the Poisson equation wxx (x) = f (x) (without relying upon
distribution theory).
90 CHAPTER 2. INTRODUCTION TO SOBOLEV SPACES

Problem 2.35. Let T ∈ S 0 (Rn ) and f ∈ S(Rn ) . Show that the Leibniz rule
∂ ∂T
for distributional derivatives holds; that is, show that (f T ) = f +
∂xi ∂xi
∂f
T in the sense of distribution.
∂xi
2 2
Problem 2.36. Let f (x) = e−s|x| and g(x) = e−t|x| . Find the Fourier
transform of f (and g) and use the inversion formula to compute f ∗ g .

Problem 2.37. Let dr denote the map given by dr f (x) = f (rx) . Show that

F(dr f ) = r−n d1/r F(f ) .

Problem 2.38. Show that a function f ∈ L2 (Rn ) is real if and only if


fˆ(−ξ) = fˆ(ξ) .
2
Problem 2.39. Find the Fourier transform of the function f (x) = xetx
for t < 0 .

Problem 2.40. Find the Fourier transform of 1(−a,a) , the characteristic


(indicator) function of the set (−a, a) .

Problem 2.41. Let f (x) = 1(0,∞) (x)e−tx , that is,

e−tx if x > 0 ,

f (x) =
0 if x ≤ 0 .

Find the Fourier transform of f for t > 0 .

Problem 2.42. Find the Fourier transform of the function f (x) = x1 |x|α ,
where x1 is the first component of x and −n − 2 < α < −2 .
Hint: Use the fact that for −n < α < 0 ,

Γ( n+α
2 ) α+ n
F(|x|α )(ξ) = 2 2 |ξ|−(α+n)
Γ(− α2 )

1 ∂
and f (x) = |x|α+2 .
α + 2 ∂x1
Problem 2.43. Let α > 0 be given. Show that the Fourier transform of the
function

1
Z
2
f (x) = tα−1 e−t e−t|x| dt
Γ(α) 0
is positive.
2.3. THE FOURIER TRANSFORM 91

Problem 2.44. Let f ∈ L1 (R) . Show that the anti-derivative of f can be


written as the convolution of f and a function ϕ ∈ L1loc (R) .

Proof. Let ϕ be the characteristic function of the set (0, ∞) , or



1 if x > 0 ,
ϕ(x) =
0 if x ≤ 0 .

Then ϕ ∈ L1loc (R) , and


Z Z x
(ϕ ∗ f )(x) = ϕ(x − y)f (y)dy = f (y)dy
R −∞

which is the anti-derivative of f . 

Problem 2.45. Let f be a continuous function with period 2π , and fˆ be


the Fourier transform of f . Show that


fˆ(ξ) =
X
( 2πfn )τ−n δ
n=−∞

in the sense of distribution, where fn is the Fourier coefficient defined by


Z 2π
1
fn = f (x)e−inx dx .
2π 0
Problem 2.46. Using Definition 2.127, compute the Fourier transform of
the function/distribution
(
x if x ≥ 0 ,
R(x) =
0 otherwise ,

by completing the following:


1
(1) Let H be the Heaviside function. Show that Ĥ(ξ) = p.v. √ +Cδ(ξ)
2πiξ
1
for some constant C , where p.v. is defined as
ξ
D 1 E
Z
ϕ(ξ)  Z − Z ∞  ϕ(ξ)
p.v. , ϕ = lim dξ = lim + dξ .
ξ →0+ R\[−,] ξ →0+ −∞  ξ

Note that the integral above always exists as long as ϕ ∈ S(R) .


1
(2) Let S(x) = H(x) − . Then S is an odd function, and show that
2
Ŝ(ξ) = −Ŝ(−ξ) .
92 CHAPTER 2. INTRODUCTION TO SOBOLEV SPACES

(3) Use (2) to determine the constant C in (1).

(4) By the definition of Fourier transform, show that hR̂, ϕi = −ihĤ, ϕ0 i ,


and as a consequence
d
R̂(ξ) = i Ĥ(ξ) .

2.4. THE SOBOLEV SPACES H S (RN ), S ∈ R 93

2.4 The Sobolev Spaces H s (Rn ), s ∈ R


2.4.1 H s (Rn ) via the Fourier Transform

The Fourier transform allows us to generalize the Hilbert spaces H k (Rn ) for
k ∈ Z+ to H s (Rn ) for all s ∈ R, and hence study functions which possess
fractional derivatives (and anti-derivatives) which are square integrable.
p
Definition 2.142. For any s ∈ Rn , let hξi = 1 + |ξ|2 , and set

H s (Rn ) = {u ∈ S 0 (Rn ) | hξis û ∈ L2 (Rn )}


= {u ∈ S 0 (Rn ) | Λs u ∈ L2 (Rn )} ,

where Λs u = F ∗ (hξis û).

The operator Λs can be thought of as a “differential operator” of order


s, and according to Rellich’s theorem, Λ−s is a compact operator, yielding
the isomorphism
H s (Rn ) = Λ−s L2 (Rn ) .

Definition 2.143. The inner-product on H s (Rn ) is given by

(u, v)H s (Rn ) = (Λs u, Λs v)L2 (Rn ) ∀u, v ∈ H s (Rn ) .

and the norm on H s (Rn ) is

kuksH s (Rn ) = (u, u)H s (Rn ) ∀u ∈ H s (Rn ) .

The completeness of H s (Rn ) with respect to the k · kH s (Rn ) ) is induced


by the completeness of L2 (Rn ).

Theorem 2.144. For s ∈ R, (H s (Rn ), k · kH s (Rn ) ) is a Hilbert space.

Example 2.145 (H 1 (Rn )). The H 1 (Rn ) in Fourier representation is exactly


the same as the that given by Definition 2.73:
Z
2
kukH 1 (Rn ) = hξi2 kû(ξ)k2 dξ
R n
Z
= (1 + |ξ|2 )kû(ξ)k2 dξ
n
ZR
= (|u(x)|2 + |Du(x)|2 )dx ,
Rn

the last equality following from the Plancheral theorem.


94 CHAPTER 2. INTRODUCTION TO SOBOLEV SPACES

1 1
Example 2.146 (H 2 (Rn )). The H 2 (Rn ) can be viewed as interpolating
between decay required for û ∈ L2 (Rn ) and û ∈ H 1 (Rn ):
Z p
1
n 2 n
H 2 (R ) = {u ∈ L (R ) | 1 + |ξ|2 |û(ξ)|2 dξ < ∞} .
Rn

Example 2.147 (H −1 (Rn )). The space H −1 (Rn ) can be heuristically de-
scribed as those distributions whose anti-derivative is in L2 (Rn ); in terms of
the Fourier representation, elements of H −1 (Rn ) possess a transforms that
can grow linearly at infinity:

|û(ξ)|2
Z
H −1 (Rn ) = {u ∈ S 0 (Rn ) | dξ < ∞} .
Rn 1 + |ξ|2

For T ∈ H −s (Rn ) and u ∈ H s (Rn ), the duality pairing is given by

hT, ui = (Λ−s T, Λs u)L2 (Rn ) ,

from which the following result follows.

Proposition 2.148. For all s ∈ R, [H s (Rn )]0 = H −s (Rn ) .

The ability to define fractional-order Sobolev spaces H s (Rn ) allows us to


refine the estimates of the trace of a function which we previously stated in
Theorem 2.93. That result, based on the Gauss-Green theorem, stated that
the trace operator was continuous from H 1 (Rn+ ) into L2 (Rn−1 ). In fact, the
1
trace operator is continuous from H 1 (Rn+ ) into H 2 (Rn−1 ).
To demonstrate the idea, we take n = 2. Given a continuous function
u : R2 → {x1 = 0}, we define the operator

T u = u(0, x2 ) .

The trace theorem asserts that we can extend T to a continuous linear map
1
from H 1 (R2 ) into H 2 (R) so that we only lose one-half of a derivative.
1
Theorem 2.149. T : H 1 (R2 ) → H 2 (R), and there is a constant C such
that
kT uk 1 ≤ CkukH 1 (R2 ) .
H 2 (R)

Before we proceed with the proof, we state a very useful result.


2.4. THE SOBOLEV SPACES H S (RN ), S ∈ R 95

Lemma 2.150. Suppose that u ∈ S(R2 ) and define f (x2 ) = u(0, x2 ). Then

1
Z
ˆ
f (ξ2 ) = √ û(ξ1 , ξ2 )dξ1 .
2π Rξ1

Proof. fˆ(ξ2 ) = √1 √1 F ∗
R R
2π R û(ξ1 , ξ2 )dξ1 if and only if f (x2 ) = 2π R û(ξ1 , ξ2 )dξ1 ,
and
1 1
Z Z Z
√ F∗ û(ξ1 , ξ2 )dξ1 = û(ξ1 , ξ2 )dξ1 eix2 ξ2 dξ2 .
2π R 2π R R
On the other hand,

1
Z Z

u(x1 , x2 ) = F [û(ξ1 , ξ2 )] = û(ξ1 , ξ2 )eix1 ξ1 +ix2 ξ2 dξ1 dξ2 ,
2π R R

so that

1
Z Z
u(0, x2 ) = F ∗ [û(ξ1 , ξ2 )] = û(ξ1 , ξ2 )eix2 ξ2 dξ1 dξ2 .
2π R R

Proof of Theorem 2.149. Suppose that u ∈ S(R2 ) and set f (x2 ) = u(0, x1 ).
According to Lemma 2.150,

1 1
Z Z
ˆ
f (ξ2 ) = √ û(ξ1 , ξ2 )dξ1 = √ û(ξ1 , ξ2 )hξi hξi−1 dξ1
2π Rξ1 2π Rξ1
Z  1 Z 1
1 2 2
2
−2
2
≤√ |û(ξ1 , ξ2 )| hξi dξ1 hξi dξ1 ,
2π R R

and hence Z Z
|f (ξ2 )| ≤ C 2 2
|û(ξ1 , ξ2 )| hξi dξ1 2
hξi−2 dξ1 .
R R
The key to this trace estimate is the explicit evaluation of the integral
−2
R
R hξi dξ1 :
  +∞
tan−1 √ ξ1 2


1
Z
1+ξ2 1
≤ π(1 + ξ22 )− 2 .

dξ1 = (2.47)
1 + ξ12 + ξ22
p
R 1 + ξ22


−∞

2 21 ˆ 2 2 2
R R
It follows that R (1 + ξ2 ) |f (ξ2 )| dξ2 ≤C R |û(ξ1 , ξ2 )| hξi dξ1 , so that inte-
gration of this inequality over the set {ξ2 ∈ R} yields the result. Using the
density of S(R2 ) in H 1 (R2 ) completes the proof. 
96 CHAPTER 2. INTRODUCTION TO SOBOLEV SPACES

The proof of the trace theorem in higher dimensions and for general
H s (Rn ) spaces, s > 12 , replacing H 1 (Rn ) proceeds in a very similar fashion;
the only difference is that the integral R hξi−2 dξ1 is replaced by Rn−1 hξi−2s dξ1 ·
R R

· · dξn−1 , and instead of obtaining an explicit anti-derivative of this integral,


an upper bound is instead found. The result is the following general trace
theorem.
1
Theorem 2.151 (The trace theorem for H s (Rn )). For s > 2, the trace
s− 12
operator T : H s (Rn ) →H (Rn−1 ) is continuous.

We can extend this result to open, bounded, C ∞ domains Ω ⊂ Rn .

Definition 2.152. Let ∂Ω denote a closed C ∞ manifold, and let {ωl }K


l=1
denote an open covering of ∂Ω, such that for each l ∈ {1, 2, ..., K}, there
exist C ∞ -class charts ϑl which satisfy

ϑl : B(0, rl ) ⊂ Rn−1 → ωl is a C ∞ diffeomorphism .

Next, for each 1 ≤ l ≤ K, let 0 ≤ ϕl ∈ C0∞ (Ul ) denote a partition of unity


so that L
P
l=1 ϕl (x) = 1 for all x ∈ ∂Ω. For all real s ≥ 0, we define

H s (∂Ω) = {u ∈ L2 (∂Ω) : kukH s (∂Ω) < ∞} ,

where for all u ∈ H s (∂Ω),


K
X
kuk2H s (∂Ω) = k(ϕl u) ◦ ϑl k2H s (Rn−1 ) .
l=1

The space (H s (∂Ω), k · kH s (∂Ω) ) is a Hilbert space by virtue of the com-


pleteness of H s (Rn−1 ); furthermore, any system of charts for ∂Ω with sub-
ordinate partition of unity will produce an equivalent norm.
1
Theorem 2.153 (The trace map on Ω). For s > 2, the trace operator
T : Ω → ∂Ω is continuous.

Proof. Let {Ul }K


l=1 denote an n-dimensional open cover of ∂Ω such that
Ul ∩ ∂Ω = ωl . Define charts θl : Vl → Ul , as in (2.17) but with each
chart being a C ∞ map, such that ϑl is equal to the restriction of θl to the
(n − 1)-dimensional ball B(0, rl ) ⊂ Rn−1 ). Also, choose a partition of unity
0 ≤ ζl ∈ C0∞ (Ul ) subordinate to the covering Ul such that ϕl = ζl |ωl .
2.4. THE SOBOLEV SPACES H S (RN ), S ∈ R 97

Then by Theorem 2.151, for s > 12 ,

K
X K
X
kuk2 1 = k(ϕl u) ◦ ϑl k2 1 ≤C k(ϕl u) ◦ ϑl k2H s (Rn ) ≤ Ckuk2H s (Ω) .
H s− 2 (∂Ω) H s− 2 (Rn−1 )
l=1 l=1

One may then ask if the trace operator T is onto; namely, given f ∈
s− 12
H (Rn−1 ) for s > 12 , does there exist a u ∈ H s (Rn ) such that f = T u?
By essentially reversing the order of the proof of Theorem 2.149, it is possible
to answer this question in the affirmative. We first consider the case that
n = 2 and s = 1.
1
Theorem 2.154. T : H 1 (R2 ) → H 2 (R) is a surjection.

Proof. With ξ = (ξ1 , ξ2 ), we define (one of many possible choices) the func-
tion u on R2 via its Fourier representation:

hξ1 i
û(ξ1 , ξ2 ) = K fˆ(ξ1 ) 2 ,
hξi

for a constant K 6= 0 to be determined shortly. To verify that kukH 1 (R1 ) ≤


kf k 1 , note that
H 2 (R)

∞ ∞ ∞ ∞
hξ1 i2
Z Z Z Z
2 2
|û(ξ1 , ξ2 )| hξi dξ1 dξ2 = K |fˆ(ξ1 )|2
dξ1 dξ2
−∞ −∞ −∞ −∞ hξi2
Z ∞ Z ∞
1
=K |fˆ(ξ1 )|2 (1 + ξ12 ) 2 2 dξ2 dξ1
−∞ −∞ 1 + ξ 1 + ξ2
≤ Ckf k2 1 ,
H 2 (R)

where we have used the estimate (2.47) for the inequality above.
It remains to prove that u(x1 , 0) = f (x1 ), but by Lemma 2.150, it suffices
that Z ∞ √
û(ξ1 , ξ2 )dξ2 = 2π fˆ(ξ1 ) .
−∞

Integrating û, we find that


Z ∞ ∞
1
q Z
ˆ 2
û(ξ1 , ξ2 )dξ2 = K f (ξ1 ) 1 + ξ1 dξ2 ≤ Kπ fˆ(ξ1 )
−∞ −∞ 1 + ξ12 + ξ22

so setting K = 2π/π completes the proof. 
98 CHAPTER 2. INTRODUCTION TO SOBOLEV SPACES

A similar construction yields the general result.


1
Theorem 2.155. For s > 12 , T : H s (Rn ) → H s− 2 (Rn−1 ) is a surjection.

By using the system of charts employed for the proof of Theorem 2.153,
we also have the surjectivity of the trace map on bounded domains.
1
Theorem 2.156. For s > 12 , T : H s (Ω) → H s− 2 (∂Ω) is a surjection.

The Fourier representation provides a very easy proof of a simple version


of the Sobolev embedding theorem.

Theorem 2.157. For s > n/2, if u ∈ H s (Rn ), then u is continuous and

max |u(x)| ≤ CkukH s (Rn ) .

Proof. By Theorem 2.118, u = F ∗ û; thus according to Hölder’s inequality


and the Riemann-Lebesgue lemma (Theorem 2.134), it suffices to show that

kûkL1 (Rn ) ≤ CkukH s (Rn ) .

But this follows from the Cauchy-Schwarz inequality since


Z Z
|û(ξ)|dξ = |û(ξ)|hξis hξi−s dξ
Rn Rn
Z  1 Z 1
2 2
2 2s −2s
≤ |û(ξ)| hξi dξ hξi dξ
Rn Rn
≤ CkukH s (Rn ) ,

the latter inequality holding whenever s > n/2. 

Hölder’s inequality can be used to prove the following

Theorem 2.158 (Interpolation inequality). Let 0 < r < t < ∞, and s =


αr + (1 − α)t for some α ∈ (0, 1). Then

kukH s (Rn ) ≤ CkukαH r (Rn ) kuk1−α


H t (Rn ) . (2.48)

Example 2.159 (Euler equation on T2 ). On some time interval [0, T ] sup-


pose that u(x, t), x ∈ T2 , t ∈ [0, T ], is a smooth solution of the Euler equa-
tions:

∂t u + (u · D)u + Dp = 0 in T2 × (0, T ] ,
div u = 0 in T2 × (0, T ] ,
2.4. THE SOBOLEV SPACES H S (RN ), S ∈ R 99

with smooth initial condition u|t=0 = u0 . Written in components, u =


(u1 , u2 ) satisfies uit + ui ,j j j + p,i = 0 for i = 1, 2, where we are using the
Einstein summation convention for summing repeated indices from 1 to 2
and where ui ,j = ∂ui /∂xj and p,i = ∂p/∂xi .
Computing the L2 (T2 ) inner-product of the Euler equations with u yields
the equality
1d
Z Z Z
2 i j i
|u(x, t)| dx + u ,j u u dx + p,i ui dx = 0 .
2 dt T2 T2 T2
| {z } | {z }
I1 I2

Notice that
1 1
Z Z
2 j
I1 = (|u| ),j u dx = |u|2 div udx = 0 ,
2 T2 2 T2
the second equality arising from integration by parts with respect to ∂/∂xj .
Integration by parts in the integral I2 shows that I2 = 0 as well, from which
d 2
the conservation law dt ku(·, t)kL2 (T2 ) follows.
To estimate the rate of change of higher-order Sobolev norms of u relies
on the use of the Sobolev embedding theorem. In particular, we claim that
on a short enough time interval [0, T ], we have the inequality
d
ku(·, t)k2H 3 (T2 ) ≤ Cku(·, t)k3H 3 (T2 ) (2.49)
dt
from which it follows that ku(·, t)k2H 3 (T2 ) ≤ M for some constant M < ∞.
To prove (2.49), we compute the H 3 (T2 ) inner-product of the Euler equa-
tions with u:
1d X Z X Z
ku(·, t)k2H 3 (T2 ) + Dα ui ,j uj Dα ui dx+ Dα p,i Dα ui dx = 0 .
2 dt T2 T2
|α|≤3 |α|≤3

The third integral vanishes by integration by parts and the fact that Dα div u =
0; thus, we focus on the nonlinearity, and in particular, on the highest-order
derivatives |α| = 3, and use D3 to denote all third-order partial derivatives,
as well as the notation l.o.t. for lower-order terms. We see that
Z Z
3 i j 3 i
D (u ,j u )D u dx = D3 ui ,j uj D3 ui dx
T2 T2
| {z }
K1
Z Z
i 3 j 3 i
+ u ,j D u D u dx + l. o. t. dx .
T2 T2
| {z }
K2
100 CHAPTER 2. INTRODUCTION TO SOBOLEV SPACES

l. o. t. dx ≤ Ckuk3H 3 (T2 ) , so
R
By definition of being lower-order terms, T2
it remains to estimate the integrals K1 and K2 . But the integral K1 vanishes
by the same argument that proved I1 = 0. On the other hand, the integral
K2 is estimated by Hölder’s inequality:

|K2 | ≤ kui ,j kL∞ (T2 ) kD3 uj kH 3 (T2 ) kD3 ui kH 3 (T2 ) .

Thanks to the Sobolev embedding theorem, for s = 2 (s needs only to be


greater than 1),

kui ,j kL∞ (T2 ) ≤ Ckui ,j kH 2 (T2 ) ≤ kukH 3 (T2 ) ,

from which it follows that K2 ≤ Ckuk3H 3 (T2 ) , and this proves the claim.
Note well, that it is the Sobolev embedding theorem that requires the use
of the space H 3 (T2 ) for this analysis; for example, it would not have been
possible to establish the inequality (2.49) with the H 2 (T2 ) norm replacing
the H 3 (T2 ) norm.

2.4.2 Fractional-order Sobolev spaces via difference quotient


norms
The case that s > 0

Lemma 2.160. For 0 < s < 1, u ∈ H s (Rn ) is equivalent to

|u(x) − u(y)|2
ZZ
2 n
u ∈ L (R ) , dxdy < ∞ .
Rn ×Rn |x − y|n+2s

Proof. The Fourier transform shows that for h ∈ Rn ,

h·ξ
Z Z Z
2 ih·ξ 2 2
|u(x + h) − u(x)| dx = |e − 1| |û(ξ)| dξ = sin2 |û(ξ)|2 dξ .
Rn Rn Rn 2

It follows that

|u(x) − u(y)|2 sin2 h·ξ


ZZ ZZ
dxdy = 2
|û(ξ)|2 dξdh
Rn ×Rn |x − y|n+2s n
R ×R n |h| n+2s
Z h Z sin2 h·ξ i
2 2
= |û(ξ)| n+2s
dh dξ
Rn Rn |h|
Z h Z sin2 (z · ξ ) i
−2s 2s 2 |ξ|
(letting h = 2|ξ|
−1
z) = 2 |ξ| |û(ξ)| n+2s
dz dξ .
R n Rn |z|
2.4. THE SOBOLEV SPACES H S (RN ), S ∈ R 101

As the integral inside of the square brackets is rotationally invariant, it is


independent of the direction of ξ/|ξ|; as such we set ξ/|ξ| = e1 and let
z1 = z · e1 denote the first component of the vector z. It follows that

|u(x) − u(y)|2
ZZ Z
dxdy = C |ξ|2s |û(ξ)|2 dξ ,
Rn ×Rn |x − y|n+2s Rn

sin2 z1
Z
where C = dz < ∞ . 
Rn |z|n+2s

Corollary 2.161. For 0 < s < 1,


h ZZ |u(x) − u(y)|2 i1
2
kukH s (Rn ) = kukL2 (Rn ) + dxdy
Rn ×Rn |x − y|n+2s

is an equivalent norm on H s (Rn ).

For real s ≥ 0, u ∈ H s (Rn ) if and only if Dα u ∈ L2 (Rn ) for all |α| ≤ [s]
(where [s] denotes the greatest integer that is not bigger than s), and

|Dα u(x) − Dα u(y)|2


ZZ
dxdy < ∞
Rn ×Rn |x − y|n+2(s−[s])

for all |α| = [s]. Moreover, an equivalent norm on H s (Rn ) is given by

|Dα u(x) − Dα u(y)|2 i1


h X ZZ
α 2
kukH s (Rn ) = kD uk2L2 (Ω(Rn ) + n+2(s−[s])
dxdy .
Rn ×Rn |x − y|
|α|≤[s]

If u ∈ H k (Rn ), k ∈ N, and ϕ ∈ S(Rn ), application of the product


rule shows that ϕu ∈ H k (Rn ). When s 6∈ N, however, the product rule
is not directly applicable and we must rely on other means to show that
ϕu ∈ H s (Rn ).

Lemma 2.162. Suppose that u ∈ H s (Rn ) for some s ≥ 0 and ϕ ∈ S(Rn ).


Then ϕu ∈ H s (Rn ).

Proof. We first conisder the case that 0 ≤ s < 1.


By Corollary 2.160, since ϕu is clearly an L2 (Rn )-function, it suffices to
show that

|(ϕu)(x) − (ϕu)(y)|2
ZZ
dxdy < ∞ .
Rn ×Rn |x − y|n+2s
102 CHAPTER 2. INTRODUCTION TO SOBOLEV SPACES

Since |(ϕu)(x) − (ϕu)(y)| ≤ |ϕ(x) − ϕ(y)||u(x)| + |u(x) − u(y)||ϕ(y)|,


|(ϕu)(x) − (ϕu)(y)|2
ZZ
dxdy
Rn ×Rn |x − y|n+2s
|ϕ(x) − ϕ(y)|2 |u(x)|2 + |u(x) − u(y)|2 |ϕ(y)|2
ZZ
≤2 dxdy
Rn ×Rn |x − y|n+2s
|ϕ(x) − ϕ(y)|2 |u(x)|2 |u(x) − u(y)|2
ZZ ZZ
≤2 dxdy + 2kϕk ∞
L (R )n dxdy .
Rn ×Rn |x − y|n+2s Rn ×Rn |x − y|
n+2s
| {z } | {z }
I1 I2

Since u ∈ H s (Rn ), I2 < ∞. On the other hand,


hZ Z Z Z i |ϕ(x) − ϕ(y)|2 |u(x)|2
I1 = + dxdy .
Rn |x−y|≤1 Rn |x−y|≥1 |x − y|n+2s

For the integral over |x − y| ≤ 1, since ϕ ∈ S(Rn ), |ϕ(x) − ϕ(y)| ≤ C|x − y|


for some constant C. Therefore,
|ϕ(x) − ϕ(y)|2 |u(x)|2
Z Z Z Z
dxdy ≤ C |x − y|2−n−2s |u(x)|2 dxdy
Rn |x−y|≤1 |x − y|n+2s Rn |x−y|≤1
Z Z
≤C |z|2−n−2s dz |u(x)|2 dx < ∞ if s < 1 .
|z|≤1 Rn

For the remaining integral,


|ϕ(x) − ϕ(y)|2 |u(x)|2
Z Z Z Z
2
dxdy ≤ 4kϕkL∞ (Rn ) |x − y|−n−2s |u(x)|2 dxdy
Rn |x−y|≥1 |x − y|n+2s Rn |x−y|≤1
Z Z n
2 −n−2s
≤ 4kϕkL∞ (Rn ) |z| dz |u(x)|2 < ∞ if s > 0 .
|z|≥1 R
The general case of s ≥ 0 can be proved in a similar fashion, and we
leave the details to the reader. 

The case that s < 0

For s < 0, we define the space H s (Rn ) to be the dual space of H −s (Rn ) with
the corresponding dual space norm (or operator norm) defined by
hu, vi
kukH s (Rn ) = sup = sup hu, vi . (2.50)
v∈H −s (Rn ) kvkH −s (Rn ) kvkH −s (Rn ) =1

The norm defined in (2.50) is equivalent to


hZ i1
2
kukH s (Rn ) = (1 + |ξ|2s )|û(ξ)|2 dξ .
Rn
2.5. FRACTIONAL-ORDER SOBOLEV SPACES ON DOMAINS WITH BOUNDARY103

2.5 Fractional-order Sobolev spaces on domains


with boundary
2.5.1 The space H s (Rn+ )

Let Rn+ = Rn−1 × R+ denote the upper half space of Rn .

The case s = k ∈ N

The space H k (Rn+ ) is the collection of all L2 (Rn+ )-functions so that the α-th
weak derivatives belong to L2 (Rn+ ) for all |α| ≤ k, that is,
n o
H k (Rn+ ) = u ∈ L2 (Rn+ ) Dα u ∈ L2 (Rn+ ) ∀ |α| ≤ k

with norm
X
kuk2H k (Rn ) = kDα uk2L2 (Rn ) . (2.51)
+ +
|α|≤k

Note that we are not able to directly use the Fourier transform to define the
H k (Rn+ ).

Definition 2.163 (Extension operator E). Fix N ∈ N. Let (a1 , · · · , aN )


solve
N
X
(−j)` aj = 1 , ` = 0, · · · , N − 1 .
j=1

We denote by E : C(Rn+ ) → C(Rn ) the function

u(x) xn ≥ 0 ,

 if
N
(Eu)(x) = (2.52)
aj u(x0 , −jxn )
P
 if xn < 0 .
j=1

Note that the the coefficients aj solve a linear system of N equations for
N unknowns which is always solvable since the determinant never vanishes.

Theorem 2.164 (Sobolev extension theorem). The operator E has a con-


tinuous extension to an operator E : H k (Rn+ ) → H k (Rn ), k ≤ N − 1, with
N defined in (2.52).
104 CHAPTER 2. INTRODUCTION TO SOBOLEV SPACES

Proof. We must show that all derivatives of u of order not bigger than N − 1
are continuous at xn = 0. We compute Dx` n Eu:

Dx` n u(x) if xn > 0 ,




Dx` n (Eu)(x) = N
(−j)` aj (Dx` n u)(x0 , −jxn )
P
 if xn < 0 .
j=1

By the definition of aj , lim Dx` 1 (Eu)(x) = lim Dx` 1 (Eu)(x). So Eu ∈


x1 →0+ x1 →0−
H k (Rn ). Finally, the continuity of E is concluded by the following inequal-
ity:

kEukH k (Rn ) ≤ CkukH k (Rn+ ) .

Lemma 2.165. For k ∈ N, each u ∈ H k (Rn+ ) is the restriction of some


w ∈ H k (Rn ) to Rn+ , that is, u = w|Rn+ .

Proof. We define the restriction map % : H k (Rn ) → H k (Rn+ ). By Theorem


2.164, the restriction map is onto, since %E = Id on H k (Rn+ ). 

The case s 6∈ N

Next, suppose that N − 2 < s < N − 1 for some N ∈ N given in (2.52), and
let E continue to denote the Sobolev extension operator.
We define the space H s (Rn+ ) as the restriction of H s (Rn ) to Rn+ with
norm

kukH s (Rn+ ) ≡ kEukH s (Rn ) . (2.53)

When s = k ∈ N, it may not be immediately clear that the H s (Rn+ )-norm


defined by (2.53) is equivalent to the H k (Rn+ )-norm defined by (2.51). Let
k · k1 be the norm defined by (2.51) and k · k2 be the norm defined by (2.53).
It is clear that kuk1 ≤ kuk2 , and by the continuity of E, kuk2 ≤ Ckuk1 ;
therefore, k · k1 and k · k2 are equivalent if s ∈ N.

2.5.2 The Sobolev space H s (Ω)

We can now define the Sobolev spaces H s (Ω) for any open and bounded
domain Ω ⊂ Rn with smooth boundary ∂Ω.
2.5. FRACTIONAL-ORDER SOBOLEV SPACES ON DOMAINS WITH BOUNDARY105

Definition 2.166 (Smoothness of the boundary). We say ∂Ω is C k if for


each point x0 ∈ ∂Ω there exist r > 0 and a C k -function γ : Rn−1 → R such
that - upon relabeling and reorienting the coordinates axes if necessary - we
have

Ω ∩ B(x0 , r) = {x ∈ B(x0 , r) | xn > γ(x1 , · · · , xn−1 )} .

∂Ω is C ∞ if ∂Ω is C k for all k ∈ N, and Ω is said to have smooth boundary


if ∂Ω is C ∞ .

Definition 2.167 (Partition of unity). Let X be a topological space. A


partition of unity is a collection of continuous functions {χj : X → [0, 1]}
P
such that j χi (x) = 1 for all x ∈ X. A partition of unity is locally finite
if each x in X is contained in an open set on which only a finite number of
χj are non-zero. A partition of unity is subordinate to an open cover {Uj }
of X if each χi is zero on the complement of Uj .

For a domain Ω with smooth boundary, we may assume that there exist
x1 , · · · xN ∈ ∂Ω, r1 , · · · rN > 0, γj ∈ C ∞ such that, upon relabeling and
reorienting the coordinates axes if necessary,

Ω ∩ Uj = {x ∈ Uj | xn > γj (x1 , · · · , xn−1 )} where Uj = B(xj , rj ) ,


SN
and Ω ⊆ j=0 Uj , and {χj }N
j=0 is a partition of unity subordinate to the
open cover {Uj } such that χj ∈ Cc∞ (Uj ), and the function ψj defined by

ψj (x) = (x1 , · · · , xn−1 , γj (x1 , · · · , xn−1 ) + xn ) .

is a diffeomorphism between a small neighborhood Vj of Rn and supp(χj ).


Let v0 = χ0 u and vj = (χj u) ◦ ψj . Then v0 can be considered as a
function defined on Rn , and vj can be considered as a function defined on
Rn+ . We then have the following definition.

Definition 2.168. The space H s (Ω) for s > 0 is the collection of all mea-
surable functions u such that χ0 u ∈ H s (Rn ) and (χj u) ◦ ψj ∈ H s (Rn+ ). The
H s (Ω)-norm is defined by
h N
X i1/2
kukH s (Ω) = kχ0 uk2H s (Rn ) + k(χj u) ◦ ψj k2H s (Rn ) .
+
j=1
106 CHAPTER 2. INTRODUCTION TO SOBOLEV SPACES

Theorem 2.169 (Extension). Let Ω be a bounded, smooth domain. For


any open set V such that Ω ⊂⊂ U , there exists a bounded linear operator
E : H s (Ω) → H s (Rn ) such that

(i) Eu = u a.e. in Ω ,

(ii) Eu has support within V ,

(iii) kEukH s (Rn ) ≤ CkukH s (Ω) , where the constant C depends only on s, Ω
and V .

Proof. Define
N
X √ h √ i
Eu = χ0 u + χj E[( χj u) ◦ ψj ] ◦ ψj−1 ,
j=1

where E : H k (Rn+ ) → H k (Rn ) is the continuous extension defined by (2.52)


for some k ≥ s. One more constraint, supp(χj ) ⊆ V , must be imposed on
χj for all j because of (ii), while this constraint is easily satisfied if we let
rj ≤ dist(Ω, ∂U ) for all j. 

2.6 The Sobolev Spaces H s (Tn ), s ∈ R


2.6.1 The Fourier Series: Revisited

Definition 2.170. For u ∈ L1 (Tn ), define


Z
Fu(k) = ûk = (2π)−n e−ik·x u(x)dx ,
Tn
and
X
F ∗ û(x) = ûk eik·x .
k∈Zn
Note that F : L1 (Tn ) → l∞ (Zn ). If u is sufficiently smooth, then inte-
gration by parts yields

F(Dα u) = −(−i)|α| k α ûk , k α = k1α1 · · · knαn .

Example 2.171. Suppose that u ∈ C 1 (Tn ). Then for j ∈ {1, ..., n},
 
∂u ∂u −ik·x
Z
−n
F (k) = (2π) e dx
∂xj Tn ∂xj
Z
= −(2π)−n u(x) (−ikj ) e−ik·x dx
Tn
= ikj ûk .
2.6. THE SOBOLEV SPACES H S (TN ), S ∈ R 107

Note that Tn is a closed manifold without boundary; alternatively, one may


identify Tn with the [0, 1]n with periodic boundary conditions, i.e., with op-
posite faces identified.

Definition 2.172. Let s = S(Zn ) denote the space of rapidly decreasing


functions û on Zn such that for each N ∈ N,

pN (u) = sup hkiN |ûk | < ∞ ,


k∈Zn
p
where hki = 1 + |k|2 .

Then
F : C ∞ (Tn ) → s , F ∗ : s → C ∞ (Tn ) ,

and F ∗ F = Id on C ∞ (Tn ) and FF ∗ = Id on s. These properties smoothly


extend to the Hilbert space setting:
F : L2 (Tn ) → l2 F ∗ : l2 → L2 (Tn )
F ∗ F = Id on L2 (Tn ) FF ∗ = Id on l2 .

Definition 2.173. The inner-products on L2 (Tn ) and l2 are


Z
n
(u, v)L2 (Tn ) = (2π)− 2 u(x)v(x)dx
Tn

and
X
(û, v̂)l2 = ûk v̂k ,
k∈Zn
respectively.

Parseval’s identity shows that kukL2 (Tn ) = kûkl2 .

Definition 2.174. We set

D0 (Tn ) = [C ∞ (Tn )]0 and s0 = [s]0 .

The space D0 (Tn ) is termed the space of periodic distributions.

In the same manner that we extended the Fourier transform from S(Rn )
to S 0 (Rn ) by duality, we may produce a similar extension to the periodic
distributions:
F : D0 (Tn ) → s0 F ∗ : s0 → D0 (Tn )
F ∗ F = Id on D0 (Tn ) FF ∗ = Id on s0 .
108 CHAPTER 2. INTRODUCTION TO SOBOLEV SPACES

Definition 2.175 (Sobolev spaces H s (Tn )). For all s ∈ R, the Hilbert
spaces H s (Tn ) are defined as follows:

H s (Tn ) = {u ∈ D0 (Tn ) | kukH s (Tn ) < ∞} ,

where the norm on H s (Tn ) is defined as


X
kuk2H s (Tn ) = |ûk |2 hki2s .
k∈Zn

The space (H s (Tn ), k · kH s (Tn ) ) is a Hilbert space, and we have that

H −s (Tn ) = [H s (Tn )]0 .

For any s ∈ R, we define the operator Λs as follows: for u ∈ D0 (Tn ),


X
Λs u(x) = |ûk |2 hkis eik·x .
k∈Zn

It follows that
H s (Tn ) = Λ−s L2 (Tn ) ,

and for r, s ∈ R,

Λs : H r (Tn ) → H r−s (Tn ) is an isomorphism .

Notice then that for any  > 0,

Λ− : H s (Tn ) → H s (Tn ) is a compact operator ,

as it is an operator-norm limit of finite-rank operators. (In particular, the


eigenvalues of Λ− tend to zero in this limit.) Hence, the inclusion map
H s+ (Tn ) ,→ H s (Tn ) is compact, and we have the following

Theorem 2.176 (Rellich’s theorem on Tn ). Suppose that a sequence uj


satisfies for s ∈ R and  > 0,

sup kuj kH s+ (Tn ) ≤ M < ∞ for a constant M 6= M (j) .

Then there exists a subsequence ujk → u in H s (Tn ).


2.6. THE SOBOLEV SPACES H S (TN ), S ∈ R 109

2.6.2 The Poisson Integral Formula and the Laplace operator

For f : S1 → R, denote by PI(f )(r, θ) the harmonic function on the unit


disk D = {x ∈ R2 : |x| < 1} with trace f :

∆ PI(f ) = 0 in D
PI(f ) = f on ∂D = S1 .

PI(f ) has an explicit representation via the Fourier series

fˆk r|k| eikθ r < 1, 0 ≤ θ < 2π ,


X
PI(f )(r, θ) = (2.54)
k∈Z

as well as the integral representation

1 − r2 f (φ)
Z
PI(f )(r, θ) = dφ r < 1, 0 ≤ θ < 2π .
2π S1 r2 − 2r cos(θ − φ) + 1
(2.55)

The dominated convergence theorem shows that if f ∈ C 0 (S1 ), then PI(f ) ∈


C ∞ (D) ∩ C 0 (D).
1
Theorem 2.177. PI extends to a continuous map from H k− 2 (S1 ) to H k (D)
for all k ∈ Z+ .

Proof. Define u = PI(f ).


1
Step 1. The case that k = 0. Assume that f ∈ H − 2 (Γ) so that

|fˆk |2 hki−1 ≤ M0 < ∞ .


X

k∈Z

Since the functions {r|k| eikθ : k ∈ Z} are orthogonal with respect to the
L2 (D) inner-product,
2
Z 2π Z 1 X
kuk2L2 (D) = fˆk r e|k| ikθ
r dr dθ


0 0
k∈Z

Z 1
|fˆk |2 |fˆk |2 (1 + |k|)−1 ≤ πkf k2
X X
≤ 2π r2|k|+1 dr = π 1 ,
0 H 2 (S1 )
k∈Z k∈Z

where we have used the monotone convergence theorem for the first inequal-
ity.
110 CHAPTER 2. INTRODUCTION TO SOBOLEV SPACES

1
Step 2. The case that k = 1. Next, suppose that f ∈ H 2 (Γ) so that

|fˆk |2 hki1 ≤ M1 < ∞ .


X

k∈Z

Since we have shown that u ∈ L2 (D), we must now prove that uθ = ∂θ u


and ur = ∂r u are both in L2 (D). Notice that by definition of the Fourier
transform and (2.54),

PI(f ) = PI(fθ ) . (2.56)
∂θ
1 1
By definition, ∂θ : H 2 (S1 ) → H − 2 (S1 ) continuously, so that for some con-
stant C,
kfθ k 1 ≤ Ckf k 1 .
H − 2 (S1 ) H 2 (S1 )

It follows from the analysis of Step 1 and (2.56) that (with u = PI(f )),

kuθ kL2 (D) ≤ Ckf k 1 .


H 2 (S1 )

Next, using the identity (2.54) notice that |rur | = |uθ |. It follows that

krur kL2 (D) ≤ Ckf k 1 . (2.57)


H 2 (S1 )

By the interior regularity of −∆ proven in Theorem 2.178, ur (r, θ) is smooth


on {r < 1}; hence the bound (2.57) implies that, in fact,

kur kL2 (D) ≤ Ckf k 1 ,


H 2 (S1 )

and hence
kukH 1 (D) ≤ Ckf k 1 ,
H 2 (S1 )
1
Step 3. The case that k ≥ 2. Since f ∈ H k− 2 (S1 ), it follows that

k∂θk f k 1 ≤ Ckf k 1
H − 2 (S1 ) H k− 2 (S1 )

and by repeated application of (2.56), we find that

kukH k (D) ≤ Ckf k 1 .


H k− 2 (S1 )


2.6. THE SOBOLEV SPACES H S (TN ), S ∈ R 111

The Hölder spaces on D are defines as follows: if u : D → R is bounded


and continuous, we write

kukC(D) := sup |u(x)| .


x∈D

For 0 < α ≤ 1, the αth -Hölder seminorm of u is


|u(x) − u(y)|
 
[u]C 0,α (D) := sup
x,y∈D,x6=y |x − y|α

and the αth -Hölder norm of u is

kukC 0,α (D) = kukC(D) + [u]C 0,α (D) .

According to Morrey’s inequality, if f ∈ H 3/2 (S 1 ) then for 0 < α < 1, f ∈


C 0,α (S1 ). Next, we use the result of Problem 2.49, together with Morrey’s
inequality (once again) and Theorem 2.90 to prove that u ∈ C 0,α (D). Let
us explain this.
We first prove the following:

f ∈ H 3/2 (S1 ) implies that f ∈ H 1/2+α (S1 ) for α ∈ (0, 1) which implies that f ∈ C 0,α (S1 ) ,

the last assertion meaning that |f (x + y) − f (x)| ≤ C|y|α .


We start with the identity

X
|f (x + y) − f (y)| = fˆk eikx (eiky − 1)


k∈Z


fˆk e (e − 1)
X ikx iky

=
k6=0
 1  1
2 2

ˆ
X X
2 1+2α   iky 2 −1−2α 
≤  |fk | hki |e − 1| hki
k6=0 k6=0
 1
2
X
iky 2 −1−2α 
= kf kH 1/2+α  |e − 1| hki .
k6=0

1
We consider |y| ≤ 2 and break the sum into two parts:

X X X
|eiky − 1|2 hki−1−2α = |eiky − 1|2 hki−1−2α + |eiky − 1|2 hki−1−2α .
k6=0 1 1
|k|≤ |y| |k|≥ |y|
112 CHAPTER 2. INTRODUCTION TO SOBOLEV SPACES

For the second sum, we use that |eiky − 1|2 ≤ 4 and employ the integral test
R∞
to see that 1/|y| r−1−2α dr ≤ C|y|2α . For the first sum, we Taylor expand
about y = 0: eiky − 1 = iky + O(y 2 ). Once again, we employ the integral
test:
X Z 1/|y|
|eiky − 1|2 hki−1−2α ≤ |y|2 + |y|2 r2 r−1−2α dr ≤ C(|y|2 + |y|2α ) .
1 1
|k|≤ |y|

Since |y| ≤ 1/2, we see that


X
|eiky − 1|2 hki−1−2α ≤ C|y|α
k6=0

as α < 1.
Next, according to Theorem 2.177, if f ∈ H 3/2 (S1 ), then u solves −∆u =
0 in D with u = f on ∂D, and kukH 2 (D) ≤ Ckf kH 3/2 (S1 ) . By Theorem 2.90,

kDukLq (D) ≤ C qkukH 2 (D) ∀q ∈ [1, ∞).

Hence, by Morrey’s inequality, we see that u ∈ C 0,1−2/q (D), and thus in


C 0,α (D) for α ∈ (0, 1).

2.6.3 Exercises

Problem 2.47. Let D := B(0, 1) ⊂ R2 and let u satisfy the Neumann


problem

∆u = 0 in D , (2.58a)
∂u
= g in ∂D := S1 . (2.58b)
∂r
If u = P I(f ) :=
P ˆ |k| eikθ , show that for f ∈ H 3/2 (S1 ),
k∈Z fk r

g = N f, (2.59)

which is the same as


ĝk = |k|fˆk .
P ˆ ikθ
N denotes the Dirichlet to Neumann map given by N f (θ) = k∈Z fk |k|e

or N f = −i ∂θ Hf = −iH ∂f∂θ , where H is the Hilbert transform, defined by
Hu(θ) = k∈Z (sgn k)ĝk eikθ .
P
2.7. REGULARITY OF THE LAPLACIAN ON Ω 113

−1 ikθ . Show
P
Problem 2.48. Define the function K(θ) = k6=0 |k| e that
K ∈ L2 (S1 ) ⊂ L1 (S1 ). Next, show that if g ∈ L2 (S1 ) and S1 g(θ)dθ =
R
0, a
solution to (2.59) is given by f (θ) = (2π)−1 S1 K(θ − φ)g(φ)dφ.
R

Problem 2.49. Consider the solution to the Neumann problem (2.58a) and
(2.58b). Show that g ∈ H 1/2 (S1 ) implies that u ∈ H 2 (D) and that
 
kuk2H 2 (D) ≤ C kgk2H 1/2 (S1 ) + kuk2L2 (D) .

2.7 Regularity of the Laplacian on Ω

We have studied the regularity properties of the Laplace operator on D =


B(0, 1) ⊂ R2 using the Poisson integral formula. These properties continue
to hold on more general open, bounded, C ∞ subsets Ω of Rn .
We revisit the Dirichlet problem

∆u = 0 in Ω , (2.60a)
u = f on ∂Ω . (2.60b)
1
Theorem 2.178. For k ∈ N, given f ∈ H k− 2 (∂Ω), there exists a unique
solution u ∈ H k (Ω) to (2.60) satisfying

kukH k (Ω) ≤ Ckf k 1 , C = C(Ω) .


H k− 2 (∂Ω)

Proof. Step 1. k = 1. We begin by converting (2.60) to a problem with ho-


mogeneous boundary conditions. Using the surjectivity of the trace operator
provided by Theorem 2.156, there exists F ∈ H 1 (Ω) such that T (F ) = f
on ∂Ω, and kF kH 1 (Ω) ≤ Ckf k 1 . Let U = u − F ; then U ∈ H 1 (Ω)
H 2 (∂Ω)
and by linearity of the trace operator, T (U ) = 0 on ∂Ω. It follows from
Theorem 2.95 that U ∈ H01 (Ω) and satisfies −∆U = ∆F in H01 (Ω); that is
h−∆U, vi = h∆F, vi for all v ∈ H01 (Ω).
According to Remark 2.106, −∆ : H01 (Ω) → H −1 (Ω) is an isomorphism,
so that ∆F ∈ H −1 (Ω); therefore, by Theorem 2.105, there exists a unique
weak solution U ∈ H01 (Ω), satisfying
Z
DU · Dv dx = h∆F, vi ∀v ∈ H01 (Ω) ,

114 CHAPTER 2. INTRODUCTION TO SOBOLEV SPACES

with
kU kH 1 (Ω) ≤ Ck∆F kH −1 (Ω) , (2.61)

and hence

u = U + F ∈ H 1 (Ω) and kukH 1 (Ω) ≤ kf k 1 .


H 2 (∂Ω)

Step 2. k = 2. Next, suppose that f ∈ H 1.5 (∂Ω). Again employing


Theorem 2.156, we obtain F ∈ H 2 (Ω) such that T (F ) = f and kF kH 2 (Ω) ≤
Ckf kH 1.5 (∂Ω) ; thus, we see that ∆F ∈ L2 (Ω) and that, in fact,
Z Z
DU · Dv dx = ∆F v dx ∀v ∈ H01 (Ω) . (2.62)
Ω Ω

We first establish interior regularity. Choose any (nonempty) open sets


Ω1 ⊂⊂ Ω2 ⊂⊂ Ω and let ζ ∈ C0∞ (Ω2 ) with 0 ≤ ζ ≤ 1 and ζ = 1 on Ω1 . Let
0 = min dist(spt(ζ), ∂Ω2 )/2. For all 0 <  < 0 , define U  (x) = η ∗ U (x)
for all x ∈ Ω2 , and set
v = −η ∗ (ζ 2 U  ,j ),j .

Then v ∈ H01 (Ω) and can be used as a test function in (2.62); thus,
Z Z
− U,i η ∗ (ζ U ,j ),ji dx = − U,i η ∗ [ζ 2 U  ,ij +2ζζ,i U  ,j ],j dx
2 
Ω Ω
Z Z
= ζ U ,ij U ,ij dx − 2 η ∗ [ζζ,i U  ,j ],j U,i dx ,
2  
Ω2 Ω

and
Z Z Z
2 
∆F v dx = − ∆F η ∗ (ζ U ,j ),j dx = − ∆F η ∗ [ζ 2 U  ,jj +2ζζ,j U  ,j ] dx .
Ω Ω2 Ω2

By Young’s inequality (Theorem 2.53),

kη ∗ [ζ 2 U  ,jj +2ζζ,j U  ,j ]kL2 (Ω2 ) ≤ kζ 2 U  ,jj +2ζζ,j U  ,j kL2 (Ω2 ) ;

hence, by the Cauchy-Young inequality with δ, Lemma 2.52, for δ > 0,


Z
∆F v dx ≤ δkζD2 U  k2L2 (Ω2 ) + Cδ [kDU  k2L2 (Ω2 ) + k∆F k2L2 (Ω) ] .

Similarly,
Z
2 η ∗ [ζζ,i U  ,j ],j U,i dx ≤ δkζD2 U  k2L2 (Ω2 ) + Cδ [kDU  k2L2 (Ω2 ) + k∆F k2L2 (Ω) ] .

2.7. REGULARITY OF THE LAPLACIAN ON Ω 115

By choosing δ < 1 and readjusting the constant Cδ , we see that

kD2 U  k2L2 (Ω1 ) ≤ kζD2 U  k2L2 (Ω2 ) ≤ Cδ [kDU  k2L2 (Ω2 ) + k∆F k2L2 (Ω) ]
≤ Cδ k∆F k2L2 (Ω) ,

the last inequality following from (2.61), and Young’s inequality.


Since the right-hand side does not depend on  > 0, there exists a sub-
sequence
0
D2 U  * W in L2 (Ω1 ) .

By Theorem 2.78, U  → U in H 1 (Ω1 ), so that W = D2 U on Ω1 . As weak


convergence is lower semi-continuous, kD2 U kL2 (Ω1 ) ≤ C k∆F kL2 (Ω) . As Ω1
2 (Ω) and that
and Ω2 are arbitrary, we have established that U ∈ Hloc

kU kH 2 ≤ Ck∆F kL2 (Ω) .


loc (Ω)

For any w ∈ H01 (Ω), set v = ζw in (2.62). Since u ∈ Hloc


2 (Ω), we may

integrate by parts to find that


Z
(−∆U − ∆F ) ζw dx = 0 ∀w ∈ H01 (Ω) .

Since w is arbitrary, and the spt(ζ) can be chosen arbitrarily close to ∂Ω, it
follows that for all x in the interior of Ω, we have that

−∆U (x) = ∆F (x) for almost every x ∈ Ω . (2.63)

We proceed to establish the regularity of U all the way to the boundary


∂Ω. Let {Ul }K
l=1 denote an open cover of Ω which intersects the boundary
∂Ω, and let {θl }K
l=1 denote a collection of charts such that

θl : B(0, rl ) → Ul is a C ∞ diffeomorphism ,
det Dθl = 1 ,
θl (B(0, rl ) ∩ {xn = 0}) → Ul ∩ ∂Ω ,
θl (B(0, rl ) ∩ {xn > 0}) → Ul ∩ Ω .

Let 0 ≤ ζl ≤ 1 in C0∞ (Ul ) denote a partition of unity subordinate to the


open covering Ul , and define the horizontal convolution operator, smoothing
functions defined on Rn in the first 1, ..., n − 1 directions, as follows:
Z
ρ ∗h F (xh , xn ) = ρ (xh − yh )F (yh , xn )dyh ,
Rn−1
116 CHAPTER 2. INTRODUCTION TO SOBOLEV SPACES

where ρ (xh ) = −(n−1) ρ(xh /), ρ the standard mollifier on Rn−1 , and xh =
(x1 , ..., xn−1 ). Let α range from 1 to n − 1, and substitute the test function

v = − ρ ∗h [(ζl ◦ θl )2 ρ ∗h (U ◦ θl ),α ],α ◦ θl−1 ∈ H01 (Ω)




into (2.62), and use the change of variables formula to obtain the identity
Z Z
k j
Ai (U ◦ θl ),k Ai (v ◦ θl ),j dx = (∆F ) ◦ θl v ◦ θl dx , (2.64)
B+ (0,rl ) B+ (0,rl )

where the C ∞ matrix A(x) = [Dθl (x)]−1 and B+ (0, rl ) = B(0, rl )∩{xn > 0}.
We define

U l = U ◦θl , and denote the horizontal convolution operator by H = ρ ∗h .

Then, with ξl = ζl ◦ θl , we can rewrite the test function as

v ◦ θl = −H [ξl2 H U l ,α ],α .

Since differentiation commutes with convolution, we have that

(v ◦ θl ),j = −H (ξl2 H U l ,jα ),α −2H (ξl ξl ,j H U l ,α ),α ,

and we can express the left-hand side of (5.54) as


Z
Aki (U ◦ θl ),k Aji (v ◦ θl ),j dx = I1 + I2 ,
B+ (0,rl )

where
Z
I1 = − Aji Aki U l ,k H (ξl2 H U l ,jα ),α dx ,
B (0,rl )
Z+
I2 = −2 Aji Aki U l ,k H (ξl ξl ,j H U l ,α ),α dx .
B+ (0,rl )

Next, we see that


Z
I1 = [H (Aji Aki U l ,k )],α (ξl2 H U l ,jα ) dx = I1a + I1b ,
B+ (0,rl )

where
Z
I1a = (Aji Aki H U l ,k ),α ξl2 H U l ,jα dx ,
B+ (0,rl )
Z
I1b = ([H , Aji Aki ]U l ,k ),α ξl2 H U l ,jα dx ,
B+ (0,rl )
2.7. REGULARITY OF THE LAPLACIAN ON Ω 117

and where

[H , Aji Aki ]U l ,k = H (Aji Aki U l ,k ) − Aji Aki H U l ,k (2.65)

denotes the commutator of the horizontal convolution operator and multi-


plication. The integral I1a produces the positive sign-definite term which
will allow us to build the global regularity of U , as well as an error term:
Z
I1a = [ξl2 Aji Aki H U l ,kα H U l ,jα +(Aji Aki ),α H U l ,k ξl2 H U l ,jα ] dx ;
B+ (0,rl )

thus, together with the right hand-side of (5.54), we see that


Z Z

ξl2 Aji Aki H U l ,kα H U l ,jα dx ≤ (Aji Aki ),α H U l ,k ξl2 H U l ,jα ] dx

B+ (0,rl ) B+ (0,rl )
Z

+ |I1b | + |I2 | + (∆F ) ◦ θl v ◦ θl dx .

B+ (0,rl )

Since each θl is a C ∞ diffeomorphism, it follows that the matrix A AT is


positive definite: there exists λ > 0 such that

λ|Y |2 ≤ Aji Aki Yj Yk ∀Y ∈ Rn .

It follows that
Z Z

λ ¯
ξl2 |∂DH l 2
 U | dx ≤

(Aji Aki ),α H U l ,k ξl2 H U l ,jα ] dx

B+ (0,rl ) B+ (0,rl )
Z

+ |I1b | + |I2 | + (∆F ) ◦ θl v ◦ θl dx ,

B+ (0,rl )

where D = (∂x1 , ..., ∂xn ) and p̄ = (∂x1 , ..., ∂xn−1 ). Application of the Cauchy-
Young inequality with δ > 0 shows that
Z Z

j k l 2 l
(A A ),α H U ,k ξl H U ,jα ] dx + |I2 | + (∆F ) ◦ θl v ◦ θl dx

B+ (0,rl ) i i

B+ (0,rl )
Z
≤δ ¯
ξl2 |∂DH l 2 2
 U | dx + Cδ k∆F kL2 (Ω) .
B+ (0,rl )

It remains to establish such an upper bound for |I1b |.


To do so, we first establish a pointwise bound for (5.45): for Ajk = Aji Aki ,
Z
j k l
[H , Ai Ai ]U ,k (x) = ρ (xh − yh )[Ajk (yh , xn ) − Ajk (xh , xn )]U l ,k (yh , xn ) dyh
B(xh ,)
118 CHAPTER 2. INTRODUCTION TO SOBOLEV SPACES

By Morrey’s inequality, |[Ajk (yh , xn ) − Ajk (xh , xn )]| ≤ CkAkW 1,∞ (B+ (0,rl )) .
Since
x − h − yh
 
1
∂xα ρ (xh − yh ) = 2 ρ0 ,
 
we see that
x − h − yh
 
1 0
  Z
j k l
∂xα [H , Ai Ai ]U ,k (x) ≤ C ρ |U l ,k (yh , xn )| dyh

B(xh ,)  
and hence by Young’s inquality,
 
j
∂xα [H , Ai Aki ]U l ,k 2 ≤ CkU kH 1 (Ω) ≤ Ck∆F kL2 (Ω) .

L (B+ (0,rl )

It follows from the Cauchy-Young inequality with δ > 0 that


Z
|I1b | ≤ δ ¯
ξl2 |∂DH l 2 2
 U | dx + Cδ k∆F kL2 (Ω) .
B+ (0,rl )

By choosing 2δ < λ, we obtain the estimate


Z
¯
ξl2 |∂DH l 2 2
 U | dx ≤ Cδ k∆F kL2 (Ω) .
B+ (0,rl )

Since the right hand-side is independent of , we find that


Z
¯
ξl2 |∂DU l 2
| dx ≤ Cδ k∆F k2L2 (Ω) . (2.66)
B+ (0,rl )

From (5.53), we know that ∆U (x) = ∆F (x) for a.e. x ∈ Ul . By the


chain-rule this means that almost everywhere in B+ (0, rl ),

−Ajk U l ,kj = Ajk ,j U l ,k +∆F ◦ θl ,

or equivalently,

−Ann Unn
l
= Ajα U l ,αj +Aβk U l ,kβ +Ajk ,j U l ,k +∆F ◦ θl .

Since Ann > 0, it follows from (5.55) that


Z
ξl2 |D2 U l |2 dx ≤ Cδ k∆F k2L2 (Ω) . (2.67)
B+ (0,rl )

Summing over l from 1 to K and combining with our interior estimates,


we have that
kukH 2 (Ω) ≤ Ck∆F kL2 (Ω) .

Step 3. k ≥ 3. At this stage, we have obtained a pointwise solution


U ∈ H 2 (Ω) ∩ H01 (Ω) to ∆U = ∆F in Ω, and ∆F ∈ H k−1 . We differentiate
this equation r times until Dr ∆F ∈ L2 (Ω), and then repeat Step 2. 
Chapter 3

Fourier Series and its


Applications

3.1 Inner product spaces and Hilbert spaces

A real/complex vector space H (that is, the associated scalar field is R/C)
is called an inner product space if to each ordered pair of x and y ∈ H, there
is associated real/complex number, so-called the inner product of x and y,
such that the following rules hold:

(1) (y, x)H = (x, y)H (the bar denotes complex conjugation);

(2) (x + y, z)H = (x, z)H + (y, z)H ;

(3) (αx, y)H = α(x, y)H if x and y ∈ H and α is a scalar;

(4) (x, x)H ≥ 0 for all x ∈ H;

(5) (x, x)H = 0 only if x = 0.

We define kxkH , the norm of the vector x, to be the non-negative square


root of (x, x). Thus kxk2H = (x, x).

Theorem 3.1 (The Schwarz inequality). For x and y ∈ H,

|(x, y)H | ≤ kxkH kykH .

Proof. Put A = kxk2H , B = |(x, y)H |, and C = kyk2H . There is a complex

119
120 CHAPTER 3. Fourier Series and its Applications

number α such that |α| = 1 and α(y, x)H = B. For any real r, we have

kx − rαyk2H = (x − rαy, x − rαy)H


= kxk2H − rα(y, x)H − rᾱ(x, y)H + r2 kyk2H
= A − 2Br + Cr2 .

Since the left-hand side is non-negative, we must have B 2 ≤ AC which gives


the desired inequality. 

Theorem 3.2 (Triangle inequality). For x and y ∈ H,

kx + ykH ≤ kxkH + kykH .

Proof. By the Schwarz inequality,

kx + yk2H = (x + y, x + y)H = kxk2H + (x, y)H + (y, x)H + kyk2H


2
≤ kxk2H + 2kxkH kykH + kyk2H = kxkH + kykH . 

Therefore, H, k · kH is a normed space (hence a metric space as well).
If (x, y)H = 0 for some x and y ∈ H, we say that x is orthogonal to y, and
sometimes write x ⊥ y. Since (x, y)H = 0 implies (y, x)H = 0, the relation
⊥ is symmetric.
If this space is complete, then H is called a Hilbert space. From now on,
we assume that H is a Hilbert space.

Theorem 3.3 (Riesz representation). If L is a continuous linear functional


on a Hilbert space H, then there is a unique y ∈ H such that

Lx = (x, y)H ∀x ∈ H.

Proof. Without loss of generality, we may assume that Lx 6= 0 for some


x ∈ H. Then ker L ( H is a closed subspace of H; hence there exists
z ∈ (ker L)⊥ with kzkH = 1. Let u = (Lx)z − (Lz)x. Then u ∈ ker L since
Lu = (Lx)(Lz) − (Lz)(Lx) = 0 . So (u, z)H = 0, or

0 = Lx(z, z)H − Lz(x, z)H ⇒ Lx = Lz(x, z)H .

Then y = (Lz)z is what we are looking for. The uniqueness of y follows


from that if (x, y) = 0 for all x ∈ H, then y = 0. 
3.1. INNER PRODUCT SPACES AND HILBERT SPACES 121

Definition 3.4 (Orthonormal sets). A set of vectors uα in a Hilbert space


H, where α runs through some index set I, is called orthonormal if it satisfies
the orthogonality relations (uα , uβ )H = 0 for all α 6= β, α ∈ I and β ∈ I,
and if it is normalized so that kuα kH = 1 for each α ∈ I. An orthonormal
set {uα }α∈I is called maximal if there is no other non-zero unit vector x ∈ H
so that (x, uα )H = 0 for all α ∈ I. In other words, {uα }α∈I is maximal if
(x, uα )H = 0 for all α ∈ I implies x = 0.

Definition 3.5 (Fourier coefficients). If {uα }α∈I is orthonormal, we asso-


ciate with each x ∈ H a real/complex function x̂ on the index set I, defined
by

x̂(α) = (x, uα )H α∈I.

The number x̂(α) sometimes is called the Fourier coefficients of x, relative


to the set {uα }α∈I .

Theorem 3.6. Suppose that {uα }α∈I is an orthonormal set in H and that
F is a finite subset of I. Let MF be the span of {uα }α∈F .

(1) If ϕ is a complex function on I that is 0 outside F , then there is a


vector y ∈ MF , namely
X
y= ϕ(α)uα
α∈F
X
that has ŷ(α) = ϕ(α) for every α ∈ I. Also, kyk2H = |ϕ(α)|2 .
α∈F
P
(2) If x ∈ H and sF (x) = x̂(α)uα , then
α∈F

kx − sF (x)kH ≤ kx − skH ∀ s ∈ MF ,

and
X
|x̂(α)|2 ≤ kxk2H . (Bessel’s inequality)
α∈F

Theorem 3.7. If {uα }α∈I is an orthonormal subset of a Hilbert space H,


then the following conditions are equivalent:
122 CHAPTER 3. Fourier Series and its Applications

(1) (x, uα )H = 0 for all α ∈ I implies that x = 0;


P
(2) x = (x, uα )H uα for all x ∈ H;
α∈I
P 2
(3) kxk2H = (x, uα )H for all x ∈ H;
α∈I

(4) The smallest closed linear subspace that contains {uα }α∈I is H;

(5) {uα }α∈I is a maximal orthonormal set.

3.2 The Hilbert space L2 (T)

A 2π-periodic function on R may be identified with a function on the circle,


or one-dimensional torus, T = R/(2πZ), which we define by identifying
points in R that differ by 2πn for some n ∈ Z. We use C(T) to denote the
space of continuous functions on R with period 2π. The space L2 (T) is the
completion of C(T) with respect to the L2 -norm
hZ i1
2
kf kL2 (T) = |f (x)|2 dx .
T

The space L2 (T) is a Hilbert space with the inner product


Z
(f, g)L2 (T) = f (x)g(x)dx .
T

The Fourier basis elements are the functions


1
ek (x) = √ eikx .

The set {ek }∞ 2


k=−∞ is an orthonormal set in L (T). The goal in this section
is to show that {ek }∞ 2
k=−∞ is maximal. In other words, for each f ∈ L (T),

∞ ∞ Z
X 1 X
f= (f, ek )L2 (T) ek = f (y)eik(x−y) dy .
2π T
k=−∞ k=−∞

3.2.1 Approximations of the identity



Definition 3.8. A family of functions ϕn ∈ C(T) n ∈ N is an approxi-
mation of the identity if

(1) ϕn (x) ≥ 0;
§3.2 The Hilbert space L2 (T) 123
Z
(2) ϕn (x)dx = 1 for every n ∈ N;
T
Z
(3) lim ϕn (x)dx = 0 for every δ > 0, here we identify T with the
n→∞ δ≤|x|≤π
interval [−π, π].

Definition 3.9 (Convolutions on T). The convolution of two continuous


function f, g : T → C is the continuous function f ∗ g : T → C defined by
the integral
Z
(f ∗ g)(x) = f (x − y)g(y) dy .
T

Note that all the conclusions from Section A.7 are still valid. In partic-
ular, we have

Theorem 3.10. If {ϕn }∞


n=1 is an approximation of the identity and f ∈
C(T), then ϕn ∗ f converges uniformly to f as n → ∞.

Proof. Without loss of generality, we may assume that f 6≡ 0. By the


definition of the convolution,
Z

(ϕn ∗ f )(x) − f (x) = ϕn (x − y)f (y)dy − f (x)
ZT

= ϕn (x − y) f (x) − f (y) dy ,
T

where we use (2) of Definition 3.8 to obtain the last equality. Now given

 > 0. Since f ∈ C(T), there exists δ > 0 such that |f (x) − f (y)| <
2
whenever |x − y| < δ. Therefore,

|(ϕn ∗ f )(x) − f (x)|


Z Z

≤ ϕn (x − y) f (x) − f (y) dy +
ϕn (x − y) f (x) − f (y) dy
|x−y|<δ δ≤|x−y|
Z Z
≤  ϕn (x − y)dy + 2 max |f | ϕn (z)dz .
T T δ≤|z|≤π

By (3) of Definition 3.8, there exists N > 0 such that if n ≥ N ,



Z
ϕn (z)dx ≤ .
δ≤|z|≤π 4 maxT |f |

Therefore, for n ≥ N , (ϕn ∗ f )(x) − f (x) ≤  . for all x ∈ T. 
124 CHAPTER 3. Fourier Series and its Applications

3.2.2 Trigonometric polynomials

Definition 3.11. A trigonometric polynomial of degree n is a finite sum of


the form
n
c0 X
f (x) = + (ck cos nx + sk sin nx) x ∈ R.
2
k=1

The collection of all trigonometric polynomial of degree n is denoted by


Pn (T).

On account of the Euler identity eiθ = cos θ+i sin θ, the expression above
can also be written in the form
n
X ck − isk
f (x) = ak eikx with ak = ,
2
k=−n

where s0 is taken to be 0.

Theorem 3.12. The trigonometric polynomials are dense in C(T) with


respect to the uniform norm.
Z
Proof. Let ϕn (x) = cn (1+cos x)n , where cn is chosen so that ϕn (x)dx = 1.
T
By the residue theorem,
(z + 1)2n
 
z + 1 n dz 1 π 2n
Z I I
(1 + cos x)n dx = 1+ = n n+1
dz = n−1
;
T S1 2z iz i2 S1 z 2 n
2n−1 (n!)2
thus cn = .
π (2n)!
Now {ϕn }∞
n=1 is clearly non-negative and satisfies (2) of Definition 3.8
for all n ∈ N. Let δ > 0 be given.
Z Z  1 + cos δ n (n!)2
ϕn (x)dx ≤ cn (1 + cos δ)n dx ≤ 22n .
δ≤|x|≤π δ≤|x|≤π 2 (2n)!
n!
By Stirling’s formula lim √ = 1,
n→∞2πnnn e−n
√ 2
2n 1 + cos δ 2πnnn e−n
Z  n
lim ϕn (x)dx ≤ lim 2 p
n→∞ δ≤|x|≤π n→∞ 2 2π(2n)(2n)2n e−2n
√  1 + cos δ n
= lim πn = 0.
n→∞ 2
So {ϕn }∞
n=1 is an approximation of the identity. By Theorem 3.10, ϕk ∗ f
converges uniformly to f if f ∈ C(T), while ϕn ∗f is a trigonometric function.

§3.2 The Hilbert space L2 (T) 125

Corollary 3.13. Given f ∈ L2 (T), define


Z π n
1 1 X ˆ
fˆ(k) = √ f (x)e−ikx dx , and sn (f, x) = √ f (k)eikx .
2π −π 2π k=−n

Then lim ksn (f, ·) − f kL2 (T) = 0, and


n→∞
Z ∞
|fˆ(k)|2 .
X
|f (x)|2 dx = (Parseval’s identity) (3.1)
T k=−∞

Proof. By Theorem 3.7 (4), it suffices to show that the smallest closed sub-
space containing {ek }∞ 2
k=−∞ is L (T). However, since every linear combina-
tion of ek can be expressed as a trigonometric polynomial, and by Theorem
3.12 the space of trigonometric polynomials is dense in C(T), we know that
the space spanned by {ek }∞
k=−∞ is dense in C(T). The implication from
uniform convergence to L2 (T)-convergence then guarantees that the space
spanned by {ek }∞ 2
k=−∞ is dense in L (T). 

Remark 3.14. Using the Euler identity, we can express sn (f, x) defined as
n
c0 X
sn (f, x) = + ck cos kx + sk sin kx ,
2
k=1

where
π π
1 1
Z Z
ck = f (x) cos kxdx , sk = f (x) sin kxdx .
π −π π −π

The corresponding Parseval’s identity is


h ∞
X i
kf k2L2 (T) = π c20 + (c2k + s2k ) .
k=1

The proof of the following lemma is left as an exercise.

Lemma 3.15. Let f, g ∈ L2 (T). Then



fˆ(k)ĝ(k).
X
(f, g)L2 (T) =
k=−∞

By Parseval’s identity (3.1), for f ∈ L2 (T),

lim |fˆ(k)| = 0 .
|k|→∞

In fact, the Fourier coefficient of a function f ∈ L1 (T) also converges to 0


which is the Riemann-Lebesgue Lemma.
126 CHAPTER 3. Fourier Series and its Applications

Lemma 3.16 (Riemann-Lebesgue). For f ∈ L1 (T), fˆ(k) → 0 as |k| → ∞.

Proof. Let f ∈ L1 (T). Given  > 0, there exists f ∈ L2 (T) such that
 
kf − f kL1 (T) < . For this , there exists N > 0 such that |fˆ (k)| <
2 2
whenever k ≥ N . Therefore, for k ≥ N ,

|fˆ(k)| ≤ |fˆ(k) − fˆ (k)| + |fˆ (k)|


Z
f (x) − f (x) ek (x)dx + |fˆ (k)|
 
=
T
≤ kf − f kL1 (T) + |fˆ (k)| < 

which implies that lim fˆ(k) = 0. 


|k|→∞

Definition 3.17 (Weak convergence). A sequence un ∈ H is said to con-


verge weakly to u ∈ H if

lim (un , g)H = (u, g)H ∀g ∈ H.


n→∞

We use the notation un * u in H to denote the weak convergence of un to


u in H.

With this definition, by the Riemann-Lebesgue Lemma we have the fol-


lowing

Theorem 3.18. The Fourier basis en converges weakly to 0 in L2 (T).

3.2.3 Fourier representation of functions on [0, π]

Any functions defined on [0, π] can be viewed as the restriction of an even/odd


function defined on [−π, π] to [0, π]. An even/odd function f in L2 (T) can
be expressed as
∞ ∞
c0 X .X
f (x) = + ck cos kx sk sin kx .
2
k=1 k=1

For a function f ∈ L2 (0, 2π) (here we identify T with [0, 2π]), g(x) = f (2x)
is a function in L2 (0, π). Since

c0 X
f (x) = + (ck cos kx + sk sin kx) ,
2
k=1
3.3. POINTWISE AND UNIFORM CONVERGENCE OF THE FOURIER SERIES127

we have

c0 X
g(x) = + (ck cos 2kx + sk sin 2kx) ,
2
k=1
where
2 2
Z Z
ck = g(x) cos kxdx , sk =
g(x) sin kxdx .
T π π T
n 1 r2 o∞ nr 2 o∞
As a consequence, √ , cos kx , sin kx are both maxi-
π π k=1
rπ k=1
r
n 1 2 2 o∞
mal orthonormal sets on L2 (0, π). So is √ , cos 2kx, sin 2kx .
r π π π k=1
n 1 2 o∞
We note that ± √ , ± cos kx is the collection of all non-trivial
π π k=1
functions with unit L2 (0, π)-norm satisfying

uxx = λu for some λ ∈ R ,


ux (0) = ux (π) = 0 ,
r
n 2 o∞
while ± sin kx is the collection of all non-trivial functions with
π k=1
2
unit L (0, π)-norm satisfying

uxx = λu for some λ ∈ R ,


u(0) = u(π) = 0 .

3.3 Pointwise and uniform convergence of the Fourier


series

Given f ∈ L2 (T), by Corollary 3.13 we know that sn (f, ·) → f in L2 (T); thus


possesses a subsequence snj (f, ·) which converges to f almost everywhere. In
this section, instead of assuming that f ∈ L2 (T), we consider f ∈ C 0,α (T),
and investigate the behavior of convergence of the Fourier series of f . We
first show that sn (f, ·) converges to f pointwise (p.w.) and then show that
the convergence is in fact uniform.
Before proceeding, we define
n
1 e−inx ei(2n+1)x − 1
 
1 X ikx
Dn (x) = e =
2π 2π eix − 1
k=−n

1 ei(n+1/2)x − e−i(n+1/2)x sin(n + 12 )x


= = .
2π eix/2 − e−ix/2 2π sin x2
128 CHAPTER 3. Fourier Series and its Applications

Then
n n
1
Z
fˆ(k)eikx =
X X
sn (f, x) = f (y)eik(x−y) dy
2π T
k=−n k=−n
Z
= f (y)Dn (x − y)dy = (Dn ∗ f )(x) ,
T

and Corollary 3.13 states that Dn ∗ f → f in L2 (T) for all f ∈ L2 (T).

Definition 3.19. The function

sin(n + 12 )x
Dn (x) = (3.2)
2π sin x2

is called the Dirichlet kernel.

3.3.1 Pointwise convergence


Z
By the construction of the Dirichlet kernel Dn , Dn (x)dx = 1 for all
T
n ∈ N. However, as one can see from (3.2), Dn attains negative values in T,
so Dn fails to be an approximation of the identity. Therefore, at first glance
we would not expect that sn (f, ·) ≡ Dn ∗ f converges to f uniformly because
one cannot apply Theorem 3.10; thus we focus on achieving the pointwise
convergence first.
Without loss of generality, we show that sn (f, 0) → 0 as n → ∞ for
otherwise we may consider g(x) = f (x+a)−f (a) and show that sn (g, 0) → 0
which implies that sn (f, a) → f (a).
By definition,

1 π sin(n + 21 )x
Z
sn (f, 0) = (Dn ∗ f )(0) = f (x) dx (3.3)
π −π 2 sin x2
1 π sin nx cos x2 1 π
Z Z
= f (x) dx + f (x) cos nxdx ;
π −π 2 sin x2 π −π

thus

1 π sin nx
Z
sn (f, 0) − f (x) dx
π −π x
1 π
Z  cos x 1 1 π
Z
2
= f (x) − sin nxdx + f (x) cos nxdx .
π −π 2 sin x2 x π −π
§3.3 Pointwise and uniform convergence of Fourier series 129

cos x2 1
Note that the function g(x) ≡ − can be made continuous at x = 0
2 sin x2 x
if
 cos x 1 x cos x2 − 2 sin x2
2
g(0) ≡ lim − = lim = 0.
x→0 2 sin x x x→0 2x sin x2
2

Therefore, by Riemann-Lebesgue Lemma,


1 π sin nx
Z
sn (f, 0) − f (x) dx → 0 as n → ∞.
π −π x

However, since f ∈ C 0,α (T) and f (0) = 0, the function f (x)/x ∈ L1 (T);
thus by Riemann-Lebesgue Lemma again,
1 π sin nx
Z
f (x) dx → 0 as n→∞
π −π x

which implies that sn (f, 0) → 0 as n → ∞. This proves the pointwise


convergence of the Fourier series of a Hölder continuous function.

3.3.2 Uniform convergence

Before proceeding to the proof of uniform convergence of the Fourier series


(of Hölder continuous functions), we make the following observation. We
consider the Fejér kernel which is the Cesàro mean of the Dirichlet kernel
defined by
n (n+1)x
1 X 1 − cos(n + 1)x 1 sin2 2
Fn (x) = Dk (x) = = .
n+1
k=0
4π(n + 1) sin2 x2 2π(n + 1) sin2 x
2
Z
We note that Fn ≥ 0 has the property that Fn (x)dx = 1. Moreover, for
T
any δ > 0,
1
Z Z
lim Fn (x)dx ≤ lim δ
dx = 0 ;
n→∞ δ<|x|≤π n→∞ δ<|x|≤π 2π(n + 1) sin2 2

hence {Fn }∞
n=1 is an approximation of the identity. Define σn (f, x) = (Fn ∗
f )(x). By Theorem 3.10, σn (f, ·) converges to f uniformly as n → ∞ if
f ∈ C(T).
However, σn (f, ·) ≡ Fn ∗ f → f uniformly as n → ∞ is still far away
from our goal: Dn ∗ f → f uniformly as n → ∞. In the following, we first
consider an easier case f ∈ C 0,1 (T); that is, f is Lipschitz continuous on
130 CHAPTER 3. Fourier Series and its Applications

T. We note that if f ∈ C 0,1 (T), then f is absolutely continuous; thus is


differentiable a.e. and satisfies the integration by parts formula
Z b x=b Z b
0
f (x)g (x)dx = f (x)g(x) − f 0 (x)g(x)dx ∀ g ∈ C 1 (T) .

a x=a a
The additional differentiability (or more precisely, the validity of the integra-
tion by parts formula) of Lipschitz continuous functions allows us to prove
the uniform convergence more easily. We have the following

Theorem 3.20. Suppose that f ∈ C 0,1 (T). Then Dn ∗ f converges to f


uniformly as n → ∞.
Z
Proof. Since Dn (x − y)dy = 1 for all x ∈ T,
T
Z

sn (f, x) − f (x) = (Dn ∗ f − f )(x) = Dn (x − y) f (y) − f (x) dy
Z T

= Dn (y) f (x + y) − f (x) dy .
T
Let L ≡ n + 1/2, then the goal is to show that
Z f (x + y) − f (x)
lim sup sin Ly dy = 0 .

L→∞ x∈T T sin y2
We break the integral into two parts, one integrating over |y| ≤ δ and
one integrating over δ < |y| ≤ π. Since f ∈ C 1 (T), |f (y + x) − f (x)| ≤
kf 0 kL∞ (T) |y|; thus
Z δ
Z f (x + y) − f (x) 0 y
sin Ly dy ≤ kf kL∞ (T) y dy ≤ Cδ . (3.4)

y
sin 2 −δ sin 2

|y|≤δ

As for the integral over δ < |y| ≤ π, we note that


Z π
f (x + y) − f (x) cos Ly f (x + y) − f (x) y=π
sin Ly y dy = −
sin 2 L sin y2

δ y=δ
Z π
cos Ly d f (x + y) − f (x)
+ dy .
δ L dy sin y2
Note that the first term on the right-hand side converges to 0 uniformly as
L → ∞. For the second term,
Z π cos Ly d f (x + y) − f (x)
dy

y
L dy sin 2

δ
Z π cos Ly f 0 (x + y) Z π cos Ly cos y f (x + y) − f (x)
2
≤ dy + dy

δ L sin y2 δ L sin2 y2
π−δ π−δ π−δ
≤ kf 0 kL∞ (T) δ
+ kf kL∞ (T) 2 δ
≤ kf kC 0,1 (T) .
L sin 2 L sin 2 L sin2 2δ
§3.3 Pointwise and uniform convergence of Fourier series 131

Therefore, combining the estimate above with (3.4), we find that


Z f (x + y) − f (x)
lim sup sup sin Ly dy ≤ Cδ ;

L→∞ x∈T T sin y2
and the conclusion follows from that δ > 0 is chosen arbitrarily. 

The uniform convergence of sn (f, ·) to f for f ∈ C 0,α (T) with α ∈ (0, 1)



requires a lot more work. The idea is to estimate f −sn (f, ·) L∞ (T) in terms
of the quantity inf kf − pkL∞ (T) . Since sn (f, ·) ∈ Pn (T), it is obvious
p∈Pn (T)
that

inf kf − pkL∞ (T) ≤ f − sn (f, ·) L∞ (T) .
p∈Pn (T)

The goal is to show the inverse inequality



f − sn (f, ·) ∞ ≤ Cn inf kf − pkL∞ (T) (3.5)
L (T) p∈Pn (T)

for some constant Cn , and pick a suitable p ∈ Pn (T) which gives a good

upper bound for f − sn (f, ·) L∞ (T) . The inverse inequality is established
via the following

Proposition 3.21. The Dirichlet kernel Dn satisfies that for all n ∈ N,


Z

Dn (x) dx ≤ 2 + log n . (3.6)
T

Proof. The validity of (3.6) for the case n = 1 is left to the reader, and we
n
X eikx
provide the proof for the case n ≥ 2 here. Recall that Dn (x) = =

k=−n
sin(n + 21 )x
. Therefore,
2π sin x2
1 sin(n + 1 )x
Z Z π Z
n
Z π
2

Dn (x) dx = 2 Dn (x) dx = 2 Dn (x) dx +

dx .

π sin x2

1
T 0 0 n

The first integral can be estimated by


Z 1
n 1 2n + 1
2 Dn (x) dx ≤ . (3.7)
0 π n
2x π
Since ≤ sin x for 0 ≤ x ≤ , the second integral can be estimated by
π 2
Z π 1 Z π
sin(n + 2 )x 1

dx ≤ dx = log π + log n . (3.8)
π sin x2 1 x

1
n n
132 CHAPTER 3. Fourier Series and its Applications

2n + 1
We then conclude (3.6) from (3.7) and (3.8) by noting that log π + ≤

2 for all n ≥ 2. 

Remark 3.22. A more subtle estimate can be done to show that


Z

Dn (x) dx ≥ c1 + c2 log n ∀n ∈ N
T

for some positive constants c1 and c2 .

With the help of Proposition 3.21, we are able to prove the inverse in-
equality (3.5). The following theorem is a direct consequence of Proposition
3.21.

Theorem 3.23. Let f be a 2π-periodic function. Then



f − sn (f, ·) ∞ ≤ (3 + log n) inf kf − pkL∞ (T) . (3.9)
L (T) p∈Pn (T)

Proof. For n ∈ N and x ∈ T,


Z

sn (f, x) ≤ Dn (y)| f (x − y) dy ≤ (2 + log n)kf kL∞ (T) .
T

Given  > 0, let p ∈ Pn (T) such that

kf − pkL∞ (T) ≤ inf kf − pkL∞ (T) + .


p∈Pn (T)

Then by the fact that sn (p, x) = p(x) if p ∈ Pn (T), we obtain that



f − sn (f, ·) ∞ ≤ f − p ∞ + p − sn (f, ·) ∞
L (T) L (T) L (T)

≤ f − p L∞ (T) + sn (f − p, ·) L∞ (T)


≤ f − p L∞ (T) + (2 + log n)kf − pkL∞ (T)
h i
≤ (3 + log n) inf kf − pkL∞ (T) +  ,
p∈Pn (T)

and (3.9) is obtained by passing  → 0. 

Having established Theorem 3.23, the study of the uniform convergence


of sn (f, ·) to f then amounts to the study of the quantity inf kf −
p∈Pn (T)
pkL∞ (T) . In Exercise Problem 3.2, the reader is asked to show that

1 + 2 log n
inf kf − pkL∞ (T) ≤ kf kC 0,1 (T) ;
p∈Pn (T) 2n
§3.3 Pointwise and uniform convergence of Fourier series 133

thus by Theorem 3.23, sn (f, ·) converges to f uniformly as n → ∞ if f ∈


C 0,1 (T), a restatement of Theorem 3.20.
The estimate of inf kf − pkL∞ (T) for f ∈ C 0,α (T) is more difficult,
p∈Pn (T)
and requires a clever choice of p. We begin with the following

Lemma 3.24. If f is a continuous function on [a, b], then for all δ1 > 0,

f (x) − f (y) ≤ 1 + δ1
 
sup sup f (x) − f (y) .
|x−y|≤δ1 δ2 |x−y|≤δ2

The proof of Lemma 3.24 is not very difficult, and is left to the readers.
Now we are in the position of prove the theorem due to D Jackson.

Theorem 3.25 (Jackson). Let f be a 2π-periodic continuous function.


Then for some constant C > 0,

inf kf − pkL∞ (T) ≤ C sup |f (x) − f (y)| .


p∈Pn (T) 1
|x−y|≤ n

Proof. Let p(x) = 1 + c1 cos x + · · · + cn cos nx be a positive trigonometric


function of degree n with coefficients {ci }ni=1 determined later. Define an
operator K on C(T) by

1
Z
Kf (x) = p(y)f (x − y)dy .
2π T

Then Kf ∈ Pn (T). Lemma 3.24 then implies

Kf (x) − f (x) ≤ 1
Z

p(y) f (x − y) − f (x) dy
2π T
Z π
1 
≤ p(y) 1 + n|y| sup f (x) − f (y) dy
2π −π |x−y|≤ n1

π
n
h Z i
= 1+ |y|p(y)dy sup f (x) − f (y) .
2π −π |x−y|≤ 1 n

π2
By Hölder’s inequality and that y 2 ≤ (1 − cos y) for y ∈ T,
2
Z π h 1 Z π i1 h 1 Z π i1
1 2 2
|y|p(y)dy ≤ y 2 p(y)dy p(y)dy
2π −π 2π −π 2π −π
hπ Z π i1 π√
2
≤ (1 − cos y)p(y)dy = 2 − c1 .
4 −π 2
134 CHAPTER 3. Fourier Series and its Applications

Therefore,
 nπ √ 
kKf − f kL∞ (T) ≤ 1 + 2 − c1 sup f (x) − f (y) .
2 |x−y|≤ 1 n


To conclude the theorem, we need to show that the number n 2 − c1 can
be made bounded by choosing p properly. Nevertheless, let
n n X
n
X (k + 1)π ikx 2 X (k + 1)π (` + 1)π i(k−`)x
p(x) = c sin e =c sin e

n+2 n+2 n+2
k=0 k=0 `=0
n n
X (k + 1)π X (k + 1)π (` + 1)π
=c sin2 + 2c sin cos(k − `)x
n+2 n+2 n+2
k=0 k, ` = 0
k 6= `
≡ 1 + c1 cos x + · · · + cn cos nx .

By comparison of coefficients of p, we find that


n n
−1
X
2 (k + 1)π 1 Xh 2(k + 1)π i
c = sin = 1 − cos
n+2 2 n+2
k=0 k=0
(2n+3)π π
n + 1 sin n+2 − sin n+2 n+2
= − π = ,
2 4 sin n+2 2

and
n n
X (k + 1)π kπ Xh π (2k + 1)π i
c1 = 2c sin sin =c cos − cos
n+2 n+2 n+2 n+2
k=1 k=1
(2n+2)π 2π i
h π sin n+2 − sin n+2
= c n cos − π
n+2 2 sin n+2
2π i
h π sin n+2
= c n cos + π
n + 2 sin n+2
π π
= c(n + 2) cos = 2 cos .
n+2 n+2
Therefore,
√  π  12 π
n 2 − c1 = n 2 − 2 cos = 2n sin
n+2 2(n + 2)
π π
= 2(n + 2) sin − 4 sin
2(n + 2) 2(n + 2)
2(n + 2) π π
=π sin − 4 sin
π 2(n + 2) 2(n + 2)
which is bounded by π + 4. 
§3.3 Pointwise and uniform convergence of Fourier series 135

Finally, since lim n−α log n = 0 for all α > 0 , we conclude the following
n→∞
Theorem 3.26. Suppose that f ∈ C 0,α (T) for some α ∈ (0, 1). Then Dn ∗f
converges to f uniformly as n → ∞.

Remark 3.27. The converse of Theorem 3.25 is the Bernstein theorem


which states that if f is a 2π-periodic function such that for some constant
C (independent of n) and α ∈ (0, 1),

inf kf − pkL∞ (T) ≤ Cn−α (3.10)


p∈Pn (T)

for all n ∈ N, then f ∈ C 0,α (T). In other words, (3.10) is an equivalent


condition to the Hölder continuity with exponent α of 2π-periodic continu-
ous functions. One way of proving the Bernstein theorem can be found in
Exercise Problem 3.4.

3.3.3 Jump discontiunity and Gibbs phenomenon

In this section, we study the convergence of the Fourier series of functions


with jump discontinuities. We show that the Fourier series evaluated at the
jump discontinuity converges to the average of the limits from the left and
the right. Moreover, the convergence of the Fourier series is never uniform
in the domain excluding these jump discontinuities due to the famous Gibbs
phenomenon: near the jump discontinuity the maximum difference between
the limit of the Fourier series and the function itself is at least 8% of the
jump. To be more precise, we have the following

Theorem 3.28. Let f : R → R be a piecewise continuously differentiable


function which is periodic with some period L > 0. Suppose that at some
point x0 the limit from the left f (x− +
0 ) and the limit from the right f (x0 ) of
the function f exist and differ by a non-zero gap a:

f (x+
0 ) − f (x0 ) = a 6= 0,

then there exists a constant c > 0, independent of f , x0 and L (in fact,


1 π sin x 1
Z
c= dx − ≈ 0.089490), such that
π 0 x 2
L
lim sn f, x0 + = f (x+
0 ) + ca , (3.11a)
n→∞ 2n
L
lim sn f, x0 − = f (x−
0 ) − ca . (3.11b)
n→∞ 2n
136 CHAPTER 3. Fourier Series and its Applications

Moreover,

f (x+
0 ) + f (x0 )
lim sn (f, x0 ) = . (3.12)
n→∞ 2
Proof. Without loss of generality, we may assume that x0 = 0 is the only
f (0+ ) + f (0− )
discontinuity of f , f (0) = , and L = 2π. Let g be a discon-
2
tinuous function defined by
 a
 (x + π) if −π ≤ x < 0 ,
 2π


g(x) = 0 if x = 0 ,

 a

 (x − π) if 0 < x ≤ π .

Then F = f + g is Lipchitz continuous on T, thus by Theorem 3.20,
f (0+ ) + f (0− )
= F (0) = lim sn (F, 0) = lim sn (f, 0) + lim sn (g, 0)
2 n→∞ n→∞ n→∞
= lim sn (f, 0) .
n→∞

This proves (3.12). Z


π
1 ia
By ĝ(k) = √ g(x)e−ikx dx = √ if k 6= 0, and ĝ(0) = 0, we
2π −π 2πk
n n
X ĝ(k) X a
find that sn (g, x) = √ eikx = − sin(kx); thus
2π πk
k=−n k=1
n n
π a kπ aX n kπ π a π sin x
X Z
sn g, =− sin =− sin →− dx .
n πk n π kπ n n π 0 x
k=1 k=1

Therefore, by the continuity of F ,


f (0+ ) + f (0− ) π π π
= lim F = lim sn f, + lim sn g,
2 n→∞ n n→∞ n n→∞ n
π  a π sin x
Z
= lim sn f, − dx ,
n→∞ n π 0 x
f (0+ ) + f (0− ) a
and (3.11a) follows from + = f (0+ ). (3.11b) can be proved
2 2
in the same fashion, and is left as an exercise. 

3.4 The Sobolev space H s (T)

Definition 3.29. For s > 0 (not necessary an integer), the Sobolev space
H s (T) consists of all functions f ∈ L2 (T) such that

|k|2s |fˆ(k)|2 < ∞ .
X

k=−∞
§3.4 The Sobolev space H s (T) 137

If f, g ∈ H s (T), then the H s (T)-inner product of f and g is defined by



(1 + |k|2 )s fˆ(k)ĝ(k)
X
(f, g)H s (T) =
n=−∞

which induces the H s (T)-norm as



(1 + |k|2 )s |fˆ(k)|2 .
X
kf k2H s (T) =
k=−∞

Example 3.30. Let the heavyside function H be defined by



1 if 0 ≤ x ≤ π ,
H(x) =
0 if −π ≤ x < 0 .
Then H ∈ L2 (T). Moreover, we have the Fourier coefficients for H :
 r
π
if k = 0,


2


Z π 
1 
Ĥ(k) = √ e−ikx dx = 0 if k is even, k 6= 0,
2π 0 

 2


if k is odd,

ik
1
hence H ∈ H s (T) if s < .
2
Theorem 3.31. Let 0 < r < t < ∞, and s = αr + (1 − α)t for some
α ∈ (0, 1). Then

kukH s (T) ≤ kukαH r (T) kuk1−α


H t (T) . (3.13)

Proof. By definition,

X
kuk2H s (T) = (1 + |k|2 )s |û(k)|2
k=−∞
X∞
= (1 + |k|2 )αr |û(k)|2α (1 + |k|2 )(1−α)t |û(k)|2(1−α) .
k=−∞

1 1
Noting that −1
+ = 1, by the Hölder inequality we find that
α (1 − α)−1

X
(1 + |k|2 )αr |û(k)|2α (1 + |k|2 )(1−α)t |û(k)|2(1−α)
k=−∞

h X ∞
iα h X i1−α
≤ (1 + |k|2 )r |û(k)|2 (1 + |k|2 )t |û(k)|2
k=−∞ k=−∞

which leads to (3.13). 


138 CHAPTER 3. Fourier Series and its Applications

3.4.1 Characterization of H 1 (T)

Definition 3.29 gives a quantitative way of describing functions in H s (T).


In this section, a qualitative point of view of H 1 (T) is provided based on
the Hahn-Banach theorem from functional analysis. Roughly speaking, a
function f ∈ H 1 (T) has weak derivatives satisfying the integration by parts
formula. We start from stating the following

Theorem 3.32 (Hahn-Banach). If Y is a linear subspace of a normed linear


space X and T : Y → R is a bounded linear functional on Y with kT k = M ,
then there is a bounded linear functional T̃ : X → R on X such that T̃
restricted to Y is equal to T and kT̃ k = M .

In other words, a bounded linear functional on a normed linear space


can be extended to a bounded linear functional on a bigger space.
Let f ∈ H 1 (T) and ϕ ∈ C 1 (T). We define a (bounded) linear functional
Tf on C 1 (T) by
Z
Tf (ϕ) = f (x)ϕ 0 (x)dx .
T

The goal is to extend Tf to a bounded linear functional T˜f defined on L2 (T).


Since the application the Hahn-Banach theorem requires that the range of
the linear function to be real, in the following discussion we will always
assume that f and ϕ are real-valued functions.
Since ϕ 0 ∈ C(T) ⊆ L2 (T), we can compute ϕ̂ 0 and obtain that ϕ̂ 0 (k) =
ik ϕ̂(k). Therefore,
Z ∞
fˆ(k)ik ϕ̂(k) ;
X
f (x)ϕ 0 (x)dx = (f, ϕ 0 )L2 (T) =
T k=−∞

hence
Z ∞
|k||fˆ(k)||ϕ̂(k)| ≤ kf kH 1 (T) kϕkL2 (T) .
X
0
f (x)ϕ (x)dx ≤

T n=−∞

The computation above shows that if f ∈ H 1 (T), Tf is a bounded linear


functional (on a subspace of L2 (T)). By the Hahn-Banach theorem, Tf can
be extended to a bounded linear functional T˜f : L2 (T) → R. By the Riesz
§3.4 The Sobolev space H s (T) 139

representation theorem, there is a function g ∈ L2 (T) such that


Z
T̃f (ϕ) = (ϕ, g)L2 (T) = ϕ(x)g(x)dx ∀ ϕ ∈ L2 (T) .
T

In particular, for ϕ ∈ C 1 (T),


Z Z
ϕ(x)g(x)dx = T̃f (ϕ) = Tf (ϕ) = f (x)ϕ 0 (x)dx .
T T
The function h = −g is called the weak derivative of f , and usually is
denoted by f 0 as well. The reason for calling h the weak derivative of f is
that if f ∈ C 1 (T), then
Z Z
0
− f (x)ϕ(x)dx = f (x)ϕ 0 (x)dx , (3.14)
T T

so h is indeed the derivative of f . Note that g ∈ L2 (T) is “the same as”


saying that f 0 ∈ L2 (T). Therefore, the space H 1 (T) consists of all L2 (T)-
functions whose weak derivatives are also L2 (T)-functions.

Remark 3.33. H 1 (T) is in fact the completion of (C 1 (T), k · kH 1 (T) ). In


other words, let f ∈ H 1 (T), then for any given  > 0, there is a function
f ∈ C 1 (T) such that

kf − f kH 1 (T) < .

Remark 3.34. The Hahn-Banach theorem does not guarantee the unique-
ness of the extension T̃f . Therefore, there might be two extensions T̃f1 and
T̃f2 mapping from L2 (T) to R that equal Tf on C 1 (T). Suppose that g1 and
g2 are the corresponding representations of T̃f1 and T̃f2 . By definition,
Z Z
ϕ(x)g1 (x)dx = T̃f1 (ϕ) = Tf (ϕ) = T̃f2 (ϕ) = ϕ(x)g2 (x)dx
T T

for all ϕ ∈ C 1 (T). Therefore, g1 = g2 a.e. in L2 (T); thus the extension T̃f
is indeed unique. The key here is that C 1 (T) is dense in L2 (T).

3.4.2 The Fourier representation of f 0 for f ∈ H 1 (T)


1
Let ϕ(x) = √ eikx in (3.14). By ϕ̂(`) = δ`k , we find that

∞ Z
ˆ ˆ
X
0
f (k) = 0
f (`)ϕ̂(`) = f 0 (x)ϕ(x)dx
`=−∞ T

ik
Z Z
=− f (x)ϕ 0 (x)dx = √ f (x)e−ikx dx = ik fˆ(k) .
T 2π T
140 CHAPTER 3. Fourier Series and its Applications

Therefore,
∞ ∞
dh 1 X ˆ i 1 X
√ f (k)eikx = f 0 (x) = √ ik fˆ(k)eikx
dx 2π k=−∞ 2π k=−∞

1 X d hˆ ikx
i
= √ f (k)e .
2π k=−∞ dx

d d
In other words, if denotes the weak differentiation operator, then
dx dx
commutes with the infinite sum (in which the convergence of the infinite
sum is understood in the L2 -sense).

Theorem 3.35 (Sobolev embedding, a simple version). If f ∈ H s (T) for


s > 1/2, then there exists f˜ ∈ C(T) so that f = f˜ almost everywhere.
Moreover, there exists a constant Cs > 0 such that

kf kL∞ (T) ≤ Cs kf kH s (T) ∀ f ∈ H s (T) . (3.15)

Proof. Let sn (f, x) be the partial sum of the Fourier series of f defined as
before. Then for n ≥ m,
1 X ˆ 1
|fˆ(k)|
X
ikx
|sn (f, x) − sm (f, x)| = √ f (k)e ≤ √
2π m<|k|≤n 2π m<|k|≤n
1 h X i1/2 h X 1 i1/2
≤√ (1 + |k|2 )s |fˆ(k)|2 .
2π m<|k|≤n (1 + |k|2 )s
m<|k|≤n

Therefore, ksn (f, ·) − sm (f, ·)kL∞ (T) → 0 as n, m → ∞, which implies that


sn (f, ·) converges uniformly; hence f˜ = lim sn (f, ·) is continuous.
n→∞
∞ i1
1 h X 1 2
The constant Cs in (3.15) can be chosen as √ 2 s
. 
2π (1 + |k| )
k=−∞

nr 1 r 2 o∞
Remark 3.36. As mentioned in Section 3.2, , cos kx is an
r π r π k=1
1 2 cos kx
orthonormal basis of L2 (0, π). Let w0 = and wk = √ . Then
π π 1 + k2
{wk }∞ 1
k=0 is an orthonormal basis of H (0, π) (see Exercise Problem 3.5).
Expand sin x in terms of this H 1 -basis, we obtain that
∞ n
2 X 4 2 X 4
sin x = − 2
cos 2kx = − lim cos 2kx ,
π π(4k − 1) π n→∞ π(4k 2 − 1)
k=1 k=1
3.5. 1-D HEAT EQUATIONS WITH PERIODIC BOUNDARY CONDITION141

where the limit is taken in the H 1 -topology, or equivalently,


n
2 X 4
lim sin x − + cos 2kx = 0.

π π(4k 2 − 1)
1
n→∞ H (0,π)
k=1

∂wk
Note that wk has the property that the derivative of wk , , vanishes at
∂x
the boundary points x = 0 and x = π for all k, but the derivative of sin x at
the boundary points does not vanish.

3.5 1-D heat equations with periodic boundary


condition

In this section, we consider the heat equation:

ut (x, t) − uxx (x, t) = f (x, t) in (0, 2π) × (0, T ) , (3.16a)


u(0, t) = u(2π, t) for all t ∈ (0, T ), (3.16b)
u(x, 0) = g(x) on (0, 2π) . (3.16c)

Condition (3.16b) is called the periodic boundary condition, which enables


us to treat solutions u(·, t) as a periodic function defined on R for all t ∈
[0, T ].
Z T Z Z T
2
We assume that |f (x, t)| dxdt = kf (·, t)k2L2 (T) dt < ∞, or f ∈
0 T 0
L2 (0, T ; L2 (T)), and g ∈ H 1 (T).

3.5.1 Formal approach

Assume that for all t ∈ [0, T ], u(·, t) ∈ L2 (T). Therefore, if dn (t) is the
Fourier coefficient of u(·, t), we can express u(x, t) as
∞ ∞
1 X X
u(x, t) = √ dk (t)eikx = dk (t)ek (x) .
2π k=−∞ k=−∞

Because of (3.16c), we must have dk (0) = ĝ(k). Moreover, for almost all
t ∈ [0, T ], f (·, t) ∈ L2 (T). Therefore,

1 X ˆ
f (x, t) = √ fk (t)eikx for almost all t ∈ (0, T ) ,
2π k=−∞
142 CHAPTER 3. Fourier Series and its Applications

where fˆk (t) is the Fourier coefficients defined by


1
Z
ˆ
fk (t) = √ f (x, t)e−ikx dx .
2π T
Suppose that we can switch the order of the differentiation and the summa-
tion, then
∞ ∞
1 X 0 1 X 2
ut (x, t) = √ dk (t)eikx , uxx (x, t) = − √ k dk (t)eikx ;
2π k=−∞ 2π k=−∞

thus by (3.16a), for almost all t ∈ [0, T ],



1 X h 0 i
√ dk (t) + k 2 dk (t) − fˆk (t) eikx = 0 . (3.17)
2π k=−∞

Since {ek }∞
k=−∞ is maximal, we find that dk (t) solves the ODE

dk0 (t) + k 2 dk (t) = fˆk (t) . (3.18)

Together with the initial condition dk (0) = ĝ(k), we find that


Z t
−k2 t 2
dk (t) = e ĝ(k) + fˆk (s)e−k (t−s) ds
0

which implies that a solution u(x, t) can be written as


∞ Z t
1 X h −k2 t 2
i
u(x, t) = √ e ĝ(k) + fˆk (s)e−k (t−s) ds eikx . (3.19)
2π k=−∞ 0

3.5.2 Rigorous approach

Before proceeding, we state a very important theorem in the study of dif-


ferential equations.

Theorem 3.37 (The Gronwall inequality). Let x(t) be a non-negative, con-


tinuous function on the interval [0, T ]. If x(t) satisfies x 0 (t) ≤ M + Cx(t)
for all t ∈ [0, T ], then
M Ct
x(t) ≤ eCt x(0) + (e − 1) ∀ t ∈ [0, T ] . (3.20)
C
Proof. Multiplying both sides of the differential inequality by the integrating
factor e−Ct , we find that
d h −Ct i
e x(t) ≤ M e−Ct .
dt
§3.5 1-D heat equations with periodic boundary condition 143

The desired inequality is then obtained by integrating the inequality above


in time from 0 to t for some t ∈ [0, T ], and the detail is left to the readers.


Corollary 3.38. Let y(t) be a non-negative, integrable function on the in-


terval [0, T ]. If y(t) satisfies
Z t
y(t) ≤ M + C y(s)ds ∀ t ∈ [0, T ] , (3.21)
0

then

y(t) ≤ M eCt ∀ t ∈ [0, T ] .


Z t
Proof. Let x(t) = y(s)ds and apply Theorem 3.37 . 
0

We truncate the Fourier presentation of g and f and find the solution to


the new problem.
n n
1 X ˆ 1 X
Let fn (x, t) = √ fk (t)eikx and gn (x) = √ ĝ(k)eikx . We
2π k=−n 2π k=−n
look for a solution un (x, t) to

unt (x, t) − unxx (x, t) = fn (x, t) in (0, 2π) × (0, T ) , (3.22a)


un (0, t) = un (2π, t) for all t ∈ (0, T ), (3.22b)
un (x, 0) = gn (x) on (0, 2π) . (3.22c)

The same procedure as the formal approach implies that


n Z t
1 X h −k2 t 2
i
un (x, t) = √ e ĝ(k) + fˆk (s)e−k (t−s) ds eikx
2π k=−n 0

is a solution to (3.22).

Energy estimates

Multiplying (3.22a) by un (x, t) and integrating over T,


Z Z Z
unt (x, t)un (x, t)dx − unxx (x, t)un (x, t)dx = fn (x, t)un (x, t)dx .
T T T
(3.23)
144 CHAPTER 3. Fourier Series and its Applications

Integrating by parts in x implies that


Z
unxx (x, t)un (x, t)dx = −kunx (·, t)k2L2 (T) .
T

1∂
Since unt (x, t)un (x, t) = |un (x, t)|2 , integrating (3.23) over time interval
2 ∂t
(0, t), we obtain
t
1
Z
kun (·, t)k2L2 (T) + kunx (·, t)k2L2 (T)
2 0
Z tZ
1
= kun (·, 0)k2L2 (T) + fn (x, s)un (x, s)dxds .
2 0 T

By Hölder’s and Young’s inequality,


Z
fn (x, s)un (x, s)dx ≤ kfn (·, s)kL2 (T) kun (·, s)kL2 (T)
T
1 1
≤ kfn (·, s)k2L2 (T) + kun (·, s)k2L2 (T) .
2 2
As a consequence,
Z t
kun (·, t)k2L2 (T) + 2 kunx (·, t)k2L2 (T)
0
h Z t i Z t
2 2
≤ kgn kL2 (T) + kfn (·, s)kL2 (T) ds + kun (·, s)k2L2 (T) ds
0 0
h Z t i Z t
≤ kgk2L2 (T) + kf (·, s)k2L2 (T) ds + kun (·, s)k2L2 (T) ds ,
0 0

where the last inequality follows from the Bessel inequality. Therefore, by
Corollary 3.38 we find that
h Z t i
kun (·, t)k2L2 (T) ≤ kgk2L2 (T) + kf (·, s)k2L2 (T) ds et ,
0

which in turn leads to that


Z t h Z T i
2 2 2
kun (·, t)kL2 (T) + 2 kunx (·, s)kL2 (T) ds ≤ C kgkL2 (T) + kf (·, s)k2L2 (T) ds
0 0

for some constant C > 0 (which depends on T ). In terms of the Fourier


coefficients,
n h
X Z t i
2
|dk (t)| + 2 |k|2 |dk (s)|2 ds ≤ M < ∞ . (3.24)
k=−n 0
§3.5 1-D heat equations with periodic boundary condition 145

Since the right-hand side of (3.24) is independent of n, we conclude that


∞ h
X Z t i
|dk (t)|2 + 2 |k|2 |dk (s)|2 ds ≤ M < ∞ (3.25)
k=−∞ 0

which is the same as saying that


Z t
2
ku(·, t)kL2 (T) + 2 kux (·, s)k2L2 (T) ds ≤ M < ∞ ,
0

1 X
where u(x, t) = √ dk (t)eikx is the limit of un (x, t).
2π k=−∞

Remark 3.39. Here we must clarify in what sense un converges to u. Since


X
kun (·, t) − u(·, t)k2H 1 (T) = (1 + |k|2 )|dk (t)|2 ,
|k|>n

by (3.24) we find that


Z T
lim kun (·, t) − u(·, t)k2H 1 (T) dt = 0 .
n→∞ 0

It is in this sense un converges to u.

In order to see u(x, t) is indeed a solution to (3.16), we prove that the


limit (under a proper sense) of (3.22) as n → ∞ is (3.16). Since fn → f in
L2 (0, T ; L2 (Ω)) and gn → g in L2 (T), it suffices to show that unt → ut and
unxx → uxx under certain sense. This amounts to show that
∞ ∞
∂ X X
dk (t)eikx = dk0 (t)eikx , (3.26a)
∂t
k=−∞ k=−∞
∞ ∞
∂2 X X
dk (t)eikx = − k 2 dk (t)eikx ; (3.26b)
∂x2
k=−∞ k=−∞

∂ ∂2
or in other words, and commute with the infinite sum.
∂t ∂x2
As discussed in Section 3.4.2, (3.26b) holds in the sense of L2 (T) if

denotes the weak differentiation in x. To be more precise, if u(x, t) =
∂x

1 X
√ dk (t)eikx ∈ H 2 (T) for almost all t ∈ (0, T ) and uxx denotes the
2π k=−∞
second weak (partial) derivatives of u in x, then

1 X 2
uxx (x, t) = − √ k dk (t)eikx ,
2π k=−∞
146 CHAPTER 3. Fourier Series and its Applications

where the infinite sum is understood in the sense of L2 (T); that is,
n
X ∞
X
2 ikx
k dk (t)e → k 2 dk (t)eikx in L2 (T) .
k=−n k=−∞

To see if u(·, t) ∈ H 2 (T) for almost all t ∈ (0, T ), we multiply (3.22a) by


unxx and the integrate over T to obtain that
Z Z
unt (x, t)unxx (x, t)dx − kunxx k2L2 (T) = fn (x, t)unxx (x, t)dx . (3.27)
T T

Integrating by parts in x implies


1d
Z Z Z
unt (x, t)unxx (x, t)dx = − untx (x, t)unx (x, t)dx = − |unx (x, t)|2 dx ;
T T 2 dt T
thus further integrating (3.27) over the time interval (0, t) leads to
Z t
1 2
kunx (·, t)kL2 (T) + kunxx (·, s)k2L2 (T) ds
2 0
Z tZ
1 2
= kunx (0, ·)kL2 (T) − fn (x, s)unxx (x, s)dxds
2 0 T
Z tZ
1
= kgnx k2L2 (T) − fn (x, s)ukxx (x, s)dxds .
2 0 T

By Hölder’s and Young’s inequality,


Z
fn (x, s)unxx (x, s)dx ≤ kfn (·, s)kL2 (T) kunxx (·, s)kL2 (T)
T
1 1
≤ kfn (·, s)k2L2 (T) + kunxx (·, s)k2L2 (T) .
2 2
Therefore, by kgnx kL2 (T) ≤ kgkH 1 (T) and kfn (·, t)kL2 (T) ≤ kf (·, t)kL2 (T) ,
Z t Z T
2 2 2
kunx (·, t)kL2 (T) + kunxx (·, s)kL2 (T) ds ≤ kgkH 1 (T) + kf (·, s)k2L2 (T) ds .
0 0

In terms of Fourier coefficients,


X n h Z t i Z T
|k|2 |dk (t)|2 + |k|4 |dk (s)|2 ds ≤ kgk2H 1 (T) + kf (·, s)k2L2 (T) ds .
k=−n 0 0

Since the right-hand side of the estimate above is independent of n, passing


n → ∞ implies that
∞ h
X Z t i Z T
|k|2 |dk (t)|2 + |k|4 |dk (s)|2 ds ≤ kgk2H 1 (T) + kf (·, s)k2L2 (T) ds .
k=−∞ 0 0
(3.28)
§3.5 1-D heat equations with periodic boundary condition 147

This shows u(·, t) ∈ H 2 (T) for almost all t ∈ (0, T ); thus (3.26b) holds for
almost all t ∈ (0, T ) provided that g ∈ H 1 (T) and f ∈ L2 (0, T ; L2 (T)).
Finally, since unxx → uxx and fn → f both in L2 (0, T ; L2 (T)), by (3.22a)

1 X 0
we also conclude that unt converges to √ dk (t)eikx in L2 (0, T ; L2 (T));
2π k=−∞
thus (3.26a) is guaranteed. To summarize, we have the following

Theorem 3.40. Suppose that f ∈ L2 (0, T ; L2 (T)), and g ∈ H 1 (T). Then


∞ Z t
1 X h −k2 t 2
i
u(x, t) = √ e ĝ(k) + fˆk (s)e−k (t−s) ds eikx (3.29)
2π k=−∞ 0

satisfies

ut (x, t) − uxx (x, t) = f (x, t) in (0, 2π) for almost all t ∈ (0, T ) ,
u(0, t) = u(2π, t) for all t ∈ (0, T ),
u(x, 0) = g(x) on (0, 2π) .

Remark 3.41. When performing the energy estimates, we do not really


make use of the exact form of dn (t) :
Z t
−k2 t 2
dk (t) = e ĝ(k) + fˆk (s)e−k (t−s) ds .
0

In other words, as long as the ODE (3.18) has a solution (which is usually
provided by the fundamental theorem of ODE), (3.25) and (3.28) can be
obtained without the knowledge of the analytic form of dk (t).

3.5.3 The special case f = 0

There are some good properties for the solution u to the heat equation (with
periodic boundary condition) when there is no external forcing. We study
these properties in this sub-section.

Maximum principle

Multiplying the heat equation ut − uxx = 0 by pu|u|p−2 and then integrating


over T, we find that
d
Z
u(·, t) p p + p(p − 1) u2x (x, t)up−2 (x, t)dx = 0 .

dt L (T)
T
148 CHAPTER 3. Fourier Series and its Applications

Integrating in time over the time interval (0, t) then implies that

u(·, t) p ≤ kgkLp (T) .
L (T)

Passing p to ∞, we find that



u(·, t) ∞ ≤ kgkL∞ (T) ∀t > 0. (3.30)
L (T)

This inequality reads that the absolute value of the solution never exceeds
the absolute value of the initial state, and is called the maximum principle
for the heat equation (with periodic boundary condition).

3.5.4 Decay estimates

Under the assumption f = 0; that is, we are in the situation that there is
no heat source in the environment,Zwe expect that the
Z solution/temperature
1
will converges to the average ḡ ≡ − g(x)dx ≡ g(x)dx as t → ∞. We
T 2π T
would like to study the convergence rate of u − ḡ.
By (3.29) with f = 0 and the Parseval identity,
2
X X
ku(·, t) − ḡk2L2 (T) = e−2k t |ĝ(k)|2 ≤ e−2t |ĝ(k)|2 ≤ e−2t kgk2L2 (T)
k6=0 k6=0

which implies that

u(·, t) − ḡ 2 ≤ e−t kgkL2 (T)



L (T)
∀t > 0.

Usually we are more interested in the case of t  1. In such a case, we


may evaluate u − ḡ in L∞ (T) and obtain that

u(·, t) − ḡ ∞ ≤ √1 −k2 t ≤ 1 kgkL1 (T) 2 2


X X
e−k (t−1) e−k

L (T)
e ĝ(k)
2π k6=0 2π
n6=0
(t ≥ 1) 1 −(t−1) X 2
≤ e kgkL1 (T) e−k ≤ Ce−t kgkL1 (T) ,

k6=0

1
where we use the fact that sup ĝ(k) ≤ √ kgkL1 (T) to conclude the in-
k∈Z 2π
equality. Moreover, suppose that g is smooth so that u is smooth, then

∂`u X 2
`
(x, t) = e−k t ĝ(k)(ik)` eikx ;
∂x
k6=0
3.6. 1-D HEAT EQUATIONS WITH DIRICHLET BOUNDARY CONDITION149

thus for all k ∈ N, similar argument implies that


∂`u X 2
(·, t) ∞ ≤ e−(t−1) kgkL1 (T) e−k |k|` ≤ C` e−t kgkL1 (T) ∀t ≥ 1.

`
∂x L (T)
k6=0

This proves the following

Theorem 3.42. Let u be the solution to the heat equation (3.16) with f = 0
and g ∈ L1 (T). Then the `-th partial derivatives of u − ḡ with respect to x
decays exponentially to zero in the uniform sense.

3.6 1-D heat equations with Dirichlet boundary


condition

In this section, we consider the following initial-boundary value problem for


the heat equation

ut (x, t) = uxx (x, t) in (0, L) × (0, ∞) , (3.31a)


u(0, t) = u(L, t) = 0 for all t , (3.31b)
u(x, 0) = f (x) ∀ x ∈ [0, L] . (3.31c)

Because of the boundary condition (3.31b), we use the orthonormal basis


nr 2 ∞
kπx o∞ X kπx
sin . Assume that u(x, t) = dk (t) sin . Then
L L k=1 L
k=1

L
π2 k2 2 kπx
Z
dk0 (t) + dk (t) = 0 ∀t > 0, dk (0) = f (x) sin dx .
L2 L 0 L

Therefore, by solving dk (t),



π 2 k2 kπx
dk (0)e− t
X
u(x, t) = L2 sin .
L
k=1

If f ∈ L2 (0, L), u defined above is a solution to (3.31).

3.7 Exercises

Problem 3.1. Prove Lemma 3.15.


150 CHAPTER 3. Fourier Series and its Applications

Problem 3.2. Let f be a 2π-periodic Lipchitz function. Show that for n ≥ 2,


1 + 2 log n
kf − Fn+1 ∗ f kL∞ (T) ≤ kf kC 0,1 (T) (3.32)
2n
and
2
f − sn (f, ·) ∞ ≤ 2π(1 + log n) kf kC 0,1 (T) .

L (T)
(3.33)
n
Hint: For (3.32), apply the estimate
nn + 1 π o
Fn (x) ≤ min ,
2π 2(n + 1)x2
in the following inequality:
h
Z δ Z −δ Z π i
f (x) − Fn+1 ∗ f (x) ≤ + + f (x + y) − f (x) Fn+1 (y)dy
−δ −π δ
π
with δ = . For (3.33), use (3.9) and note that
n+1
inf kf − pkL∞ (T) ≤ kf − Fn ∗ f kL∞ (T) .
p∈Pn (T)

Problem 3.3. A function f : T → R is said to be piecewise C 1 if there


are finitely many disjoint open intervals Ii so that f ∈ C 1 (Ii ) for all i and

Ii = T. Show that Dn ∗ f converges to f uniformly as n → ∞ on any
i
compact subset of Ii .

Problem 3.4. In this problem, we are concerned with the following

Theorem 3.43 (Bernstein). Suppose that f is a 2π-periodic function such


that for some constant C and α ∈ (0, 1),

inf kf − pkL∞ (T) ≤ Cn−α


p∈Pn (T)

for all n ∈ N. Then f ∈ C 0,α (T).

Complete the following to prove the theorem.

1. Suppose that there is p ∈ Pn (T) such that

kp 0 kL∞ (T) > n , kpkL∞ (T) < 1 , and p 0 (0) = kp 0 kL∞ (T) .
 π π
Choose γ ∈ − , such that sin(nγ) = −p(0) and cos(nγ) > 0, and
n n
π 1
define αk = γ + k+ for −n ≤ k ≤ n. Show that the function
n 2
r(x) = sin n(x − γ) − p(x) has at least one zeros in each interval
(αk , αk+1 ).
§3.7 Exercise 151

2. Let s ∈ N be such that such that 0 ∈ (αs , αs+1 ). Show that r has
at least 3 distinct zeros in (αs , αs+1 ) by noting that r0 (0) < 0 and
r(0) = 0.

3. Combining 1 and 2, show that

kp 0 kL∞ (T) ≤ nkpkL∞ (T) ∀ p ∈ Pn (T). (3.34)

4. Choose pn ∈ Pn (T) such that kf − pn k ≤ 2Cn−α for n ∈ N. Define



P
q0 = p1 , and qn = p2n − p2n−1 for n ∈ N. Show that qn = f and
n=0
the convergence is uniform.

5. Show that kqn kL∞ (T) ≤ 6C2−nα . As a consequence, show that

qn (x) − qn (y) ≤ 6Cn2n(1−α) |x − y| and qn (x) − qn (y) ≤ 12C2−nα .


6. For any x, y ∈ T with |x − y| ≤ 1, choose m ∈ N such that 2−m ≤


|x − y| ≤ 21−m . Then use the inequality

m−1
X X
f (x) − f (y) ≤ qn (x) − qn (y) + qn (x) − qn (y)
n=0 n=m

to show that f (x) − f (y) ≤ B|x − y|α for some constant B > 0.

Problem 3.5. Show that {wk }∞


k=0 defined in Remark 3.36 is an orthonormal
basis of H 1 (0, π).
Hint: Use the Parseval identity to show that {wk }∞
k=0 is a maximal or-
thonormal set of H 1 (0, π); that is, show that for all f ∈ H 1 (0, π),
Z π ∞
(f, wk )H 1 (0,π) 2 .
X
kf k2H 1 (0,π) = |f (x)|2 + |f 0 (x)|2 dx =

0 k=0

nr 2 o∞
You might need the fact that sin kx is an orthonormal basis of
π k=1
L2 (0, π).

Problem 3.6. Let f (x) = x on [−π, π]. Then f 0 (x) = 1 is certainly a



L2 (−π, π)-function. However, you may want to check that |n|2 |fˆ(n)|2 =
P
n=−∞
∞, so by “definition”, it does not seem to be an H 1 (−π, π)-function. What
is wrong with the argument?
152 CHAPTER 3. Fourier Series and its Applications

Problem 3.7. Show Remark 3.33.

Problem 3.8 (Generalized Gronwall inequality). Show that if a ∈ L1 (0, T )


is a non-negative function , and x(t) satisfies the following integral integral
inequality
Z t
x(t) ≤ M + a(s)x(s)ds .
0
Z t 
Then x(t) ≤ M exp a(s)ds for all t ∈ [0, T ]. In particular, if x satisfies
0

x0 (t) ≤ b(t) + a(t)x(t)

for some a, b ∈ L1 (0, T ) and a ≥ 0, then


   Z t 
x(t) ≤ x(0) + kbkL1 (0,T ) exp a(s)ds ∀ t ∈ [0, T ] .
0

Problem 3.9. Use Fourier series to formally solve the following initial-
boundary value problem for the wave equation

utt (x, t) = c2 uxx (x, t) in (0, 1) × R ,


u(0, t) = u(1, t) = 0 for all t ,
u(x, 0) = f (x) , ut (x, 0) = g(x) ∀ x ∈ [0, 1] .

Derive the following two conservation laws from your Fourier series solution
and directly from the PDE:
d 1h
Z i
|ut (x, t)|2 + c2 |ux (x, t)|2 dx = 0 .
dt 0
Problem 3.10. Use Fourier series to formally solve the following initial-
boundary value problem for the Schrödinger equation

iut (x, t) = −uxx (x, t) in (0, 1) × R ,


u(0, t) = u(1, t) = 0 for all t ,
u(x, 0) = f (x) ∀ x ∈ [0, 1] .

Derive the following two conservation laws from your Fourier series solution
and directly from the PDE:
d 1 d 1
Z Z
|u(x, t)|2 dx = 0 , |ux (x, t)|2 dx = 0 .
dt 0 dt 0
§3.7 Exercise 153

Problem 3.11. Try using the Fourier series to solve

ut (x, t) = uxx (x, t) in (0, π) × (0, ∞) ,


u(0, t) = ux (π, t) = 0 for all t ,
u(x, 0) = f (x) ∀ x ∈ [0, π] .

The most important task is to look for a suitable basis that fits the boundary
condition.

Problem 3.12. Let (r, θ) be the polar coordinate on R2 .

(1) Show that a harmonic function u on Ω ⊂ R2 satisfies


1 1
(rur )r + 2 uθθ = 0 r > 0.
r r
(2) For α > 0, let Ωα be the wedge given in polar coordinates (r, θ) by

Ωα = {(r, θ) | 0 < r < 1 , 0 < θ < α} .

Based on the fact that the general solution to

r2 R 00 (r) + rR 0 (r) − s2 R(r) = 0

is of the form R(r) = C1 rs + C2 r−s , use the Fourier series to find a


bounded solution to the following boundary value problem

∆u = 0 in Ωα ,
u=0 on {θ = 0, α} ,
 πθ 
u = sin on {r = 1} .
α

u=0 u = sin πθ
α

∆u = 0

u=0
P
Hint: Suppose that u(r, θ) = k Rk (r)ek (θ), where {ek } forms an
orthonormal basis of L2 (0, α) satisfying certain boundary conditions
(you have to figure out what these boundary conditions are). Solve Rk
by finding an ODE for Rk .
154 CHAPTER 3. Fourier Series and its Applications

(3) Find all α > 0 so that u ∈ C 2 (Ωα ).

Problem 3.13. Complete the following.

(1) Suppose that {en }∞ 2 ∞


n=1 is an orthonormal basis of L (0, `1 ) and {ẽm }m=1
is an orthonormal basis of L2 (0, `2 ). Show that {en (x)ẽm (y)}∞
n,m=1
forms an orthonormal basis of L2 ([0, `1 ] × [0, `2 ]).

Hint: Check the orthonormality and the maximality. For the maxi-
mality, check the Parseval identity.

(2) Solve the following PDE:

ut (x, y, t) − ∆u(x, y, t) = 0 (x, y) ∈ (0, π) × (0, π) , t > 0 ,


u(x, y, 0) = x(π − x) sin y (x, y) ∈ (0, π) × (0, π) ,
ux (0, y, t) = ux (π, y, t) = 0 y ∈ (0, π) , t > 0 ,
u(x, 0, t) = u(x, π, t) = 0 x ∈ (0, π) , t > 0 .
u=0

ut − ∆u = 0 ux = 0
ux = 0
u|t=0 = x(π − x) sin y

u=0
(3) Show that for all t ≥ 0, u from (b) satisfies
π π Z tZ π π
π6
Z Z Z
|u(x, y, t)|2 dxdy + 2 |Du(x, y, s)|2 dxdyds = .
0 0 0 0 0 60
Chapter 4

The Laplace and Poisson


Equations

Definition 4.1. On regions Ω ⊂ Rn , the Laplace operator ∆, also called the


Laplacian, is defined as
n
X ∂2
∆= .
i=1
∂x2i

Definition 4.2. A C 2 -function u satisfying ∆u = 0 is called a harmonic


function.

4.1 The fundamental solution

When Ω = Rn , the Laplace operator has radial symmetry, and we may


search for harmonic functions on Rn which depend only upon the radial
component. Letting
r = |x| ,

we look for a harmonic function u satisfying

u(x) = v(r) .

∂r xi
Since = , by the chain rule,
∂xi r

∂u dv ∂r xi ∂2u ∂ h 0 xi i x2i h 1 x2 i
= = v 0 (r) , = v (r) = v 00
(r) + v 0
(r) − 3i .
∂xi dr ∂xi r ∂x2i ∂xi r r2 r r

155
156 CHAPTER 4. The Laplace/Poisson Equations

Therefore,
n
X ∂2u n−1 0
∆u = = v 00 (r) + v (r) .
i=1
∂x2i r

Hence ∆u = 0 if and only if

n−1 0
v 00 + v = 0.
r

If we consider solutions away from r = 0 and suppose that v 0 (r) 6= 0, then

n−1
[log v 0 (r)]0 = .
r

This is a simple ordinary differential equation which we can directly integrate


to find that for r > 0,

 b log r + c (n = 2)
v(r) = b

n−2
+c (n ≥ 3) ,
r
where b and c are constants. These radially symmetric functions, harmonic
away from the origin, provide us with the singular integral kernels on Rn and
explicit representations for the solutions to the Poisson equation −∆u = f ,
at least when the forcing function f is “nice” enough.

Definition 4.3. The function


 1
 −
log |x| (n = 2)


Φ(x) = 1

 (n ≥ 3) ,
(n − 2)ωn−1 |x|n−2

defined for x ∈ Rn , x 6= 0 , is the fundamental solution of Laplace’s equation,


where ωn−1 denotes the surface area of the unit ball in Rn and is defined by

2π n/2
ωn−1 = ,
Γ(n/2)

where Γ denotes the Gamma function.

Notation. We will use that notation F,i to denote ∂F/∂xi , while F,ij
denotes ∂ 2 F/∂xi ∂xj and similarly for higher-order partial derivatives.
4.1. THE FUNDAMENTAL SOLUTION 157

A direct computation shows that


−xi
Φ,i (x) = |x|−n ,
ωn−1
1 h i
Φ,ij (x) = − |x| δij + nxi xj |x|−n−2 ,
2
ωn−1
and we have the following derivative estimates:
1
|Φ,i (x)| ≤ |x|1−n , (4.1a)
ωn−1
n
|Φ,ij (x)| ≤ |x|−n , (4.1b)
ωn−1
|Dα Φ(x)| ≤ C(n, |α|)|x|2−n−|α| , (4.1c)

where Dα and |α| are the multi-index notation defined by the following

Definition 4.4 (Multi-index). An n-dimensional multi-index is a vector


α = (α1 , · · · , αn ) of non-negative integers. |α| is defined as the sum of αk
and α! is defined as the product of αk ! , i.e.,
n
X n
Y
|α| = αk and α! = αk ! .
k=1 k=1

The differential operator Dxα is defined by


∂ α1 ∂ αn
Dxα = α1 · · · .
∂x1 ∂xαnn

When the (spatial) variable is specified, we simply use Dα to denote Dxα .

4.1.1 Uniform and Hölder Continuous Functions

For Ω ⊂ Rn open, a function u : Ω → R is Lipschitz continuous if

|u(x) − u(y)| ≤ C|x − y| ∀ x, y ∈ Ω , (4.2)

where C is a constant that depends on Ω but not on the function u itself.


The inequality (4.2) provides a uniform modulus of continuity. The standard
example of functions which are Lipschitz continuous but not differentiable is
given by u(x) = |x|. It is interesting to refine this functional framework to
be able to discern the regularity of functions u(x) = |x|α for positive α ≤ 1.
We wish to understand how “cuspy” the graph of u is near the origin, for
158 CHAPTER 4. The Laplace/Poisson Equations

example. To do so, we replace the difference quotient bound in (4.2) with


the following inequality:

|u(x) − u(y)| ≤ C|x − y|α ∀ x, y ∈ Ω . (4.3)

Functions which satisfy the inequality (4.3) are termed Hölder continuous
with exponent α.

Definition 4.5 (Continuous functions and compact support). For Ω ⊂ Rn ,


we let C 0 (Ω) denote the continuous functions on Ω, and we denote by C00 (Ω)
those functions in C 0 (Ω) with compact support contained in Ω.

Definition 4.6 (Bounded continuous functions). For Ω ⊂ Rn we set

C 0 (Ω) := {u : Ω → R | u is bounded and continuous} ,

with norm kukC 0 (Ω) = max |u(x)|. For integers k ≥ 0, we let C k (Ω) denote
x∈Ω
the functions possessing partial derivatives to all orders up to k which are
k (Ω) to denote the functions in
bounded and continuous on Ω. We use Cloc
C k (B) for all bounded balls contained in Ω.

Definition 4.7 (Hölder continuous functions). For u ∈ C 0 (Ω) and 0 < α ≤


1, we set

u ∈ C 0,α (Ω) := u ∈ C 0 (Ω) kukC 0 (Ω) + [u]C 0,α (Ω) < ∞ ,




where
|u(x) − u(y)|
[u]C 0,α (Ω) = sup .
x,y∈Ω |x − y|α
x6=y

The norm is kukC 0,α (Ω) = kukC 0 (Ω) + [u]C 0,α (Ω) .

Theorem 4.8. The space u ∈ C 0,α (Ω) endowed with the norm k · kC 0,α (Ω)
is a Banach space.

We leave the proof as an exercise for the reader.


We will denote C00,α (Ω) = C 0,α (Ω) ∩ C00 (Ω).
4.1. THE FUNDAMENTAL SOLUTION 159

4.1.2 The Poisson equation −∆u = f in Rn

Our objective, here, is to produce explicit solutions to the Poisson equation


−∆u = f in Rn . We will show that convolution between the fundamental
solution Φ and the “forcing function” f is a solution to this problem.

Definition 4.9. We set


Z
u(x) = Φ(x − y)f (y)dy (4.4)
Rn

whenever Φ(x − ·)f (·) ∈ L1 (Rn ).

Lemma 4.10. Suppose that f is bounded and integrable with compact sup-
port. Then if u is given by (4.4), u ∈ C 1 (Rn ) and for any x ∈ Rn and
i = 1, ..., n,
Z
u,i (x) = Φ,i (x − y)f (y)dy .
Rn
Proof. Since f is bounded with compact support, the integral
Z
Ii (x) ≡ Φ,i (x − y)f (y)dy
Rn
is well-defined and continuous
Z (in x). It suffices to show that it is the
derivative of u(x) ≡ Φ(x − y)f (y)dy .
Rn
Let ρ : (0, ∞) → R be a smooth, monotone increasing function such that
(
1 if z ∈ (2, ∞) ,
ρ(z) = and |ρ0 | ≤ 2 ,
0 if z ∈ (0, 1) ,
and define
Z  |x − y| 
u (x) = ρ Φ(x − y)f (y)dy .
Rn 
Note that since ρ is uniformly bounded and Φ ∈ L1 (Rn ) , by the dominated
convergence theorem, u → u uniformly as  → 0 on compact subsets.
 
Furthermore, as Dx [ρ |x−y|
 Φ(x − y)]f (y) is also integrable for  > 0, we
may differentiate under the integral to find that
∂ h  |x − y| 
Z i
u ,i (x) = ρ Φ(x − y) f (y)dy
n ∂xi 
ZR
xi − yi 0 |x − y| 

= ρ Φ(x − y)f (y)dy
Rn |x − y| 
Z  |x − y| 
+ ρ Φ,i (x − y)f (y)dy .
Rn 
160 CHAPTER 4. The Laplace/Poisson Equations

Therefore,
Z x − y  |x − y| 
i i 0
|u ,i (x) − Ii (x)| ≤ ρ Φ(x − y) f (y) dy

Rn |x − y| 

Z  |x − y| 
+ 1 − ρ Φ,i (x − y) f (y) dy .

Rn 

Note that |ρ| ≤ 1 and |ρ0 | ≤ 2 . Moreover, since spt(1 − ρ) ⊆ [0, 2] , it follows
that spt 1 − ρ · ⊆ [0, 2] . Similarly, since spt(ρ0 ) ⊆ [1, 2] , we see that


spt(ρ0 · ⊆ [, 2] . As a consequence, for the case n ≥ 3 (the case n = 2 is




left as an exercise for the reader) ,

|u ,i (x) − Ii (x)|


2 kf kL∞ (Ω) kf kL∞ (Ω)
Z Z
−n+2
≤ |x − y| dy + |x − y|1−n dy
 <|x−y|<2 ωn−1 |x−y|<2 (n − 2)ωn−1
Z 2 Z 2
kf kL∞ (Ω) h 2
Z Z i
≤ rdrdSz + drdSz
ωn−1  |z|=1  |z|=1 0
h 2 Z 2 Z 2 i
= kf kL∞ (Ω) rdr + dr → 0 as  → 0 ;
  0

hence u ,i → Ii uniformly as  → 0 .
Finally, by the uniform convergence of u to u and u ,i to Ii as  → 0 ,
we conclude that
x x
∂u
h Z i Z
u(x) = lim u (x) = lim u (x0 ) + dxi = u(x0 ) + Ii dxi ;
→0 →0 x0 ∂xi x0

thus u,i = Ii . 

Remark 4.11. Given that DΦ is integrable near the origin, it is possible


to compute the first partial derivatives of u by taking the limit of a sequence
of difference quotients of u. On the other hand, since D2 Φ is not integrable
near the origin, analysis of second partial derivatives of u require some sort
of limiting process, wherein the singular behavior at |x| = 0 is either excised
or regularized.
For example, we might consider removing a small ball near the origin,
and defining an approximation to u as follows:
Z
ũ (x) ≡ Φ(x − y)f (y)dy ,
Rn \B(x,)
4.1. THE FUNDAMENTAL SOLUTION 161

which makes ũ a differentiable function. However, as the domain of inte-


gration also depends upon x, differentiation of ũ becomes a bit complicated,
requiring a change of variables. To avoid this procedure, one alternative is
the introduction of the cut-off function ρ (introduced in the above proof ),
which has the similar affect of removing the singular region, without any
difficulties in differentiation.

Theorem 4.12. Suppose that Ω ⊂ Rn , f ∈ C00,α (Ω) with 0 < α ≤ 1 , and


suppose that u is given by (4.4). Then u ∈ C 2 (Rn ) , and for any x ∈ Rn ,
and i, j = 1, ..., n,
Z Z

u,ij (x) = Φ,ij (x − y) f (y) − f (x) dy − f (x) Φ,i (x − y)Nj dSy , (4.5)
Ω0 ∂Ω0

where Ω0 is any bounded, smooth domain containing Ω . In particular,

−∆u = f in Rn . (4.6)

Remark 4.13. Before starting the proof of Lemma 4.12, we explain why
a formula like (4.5) is well-defined. Note that as Dα Φ is not integrable
for |α| ≥ 2, ZΦ,ij (x − ·)f (·) 6∈ L1 (Rn ) even if f ∈ C0∞ (Rn ). Nevertheless,
Φ,ij (x − y) f (y) − f (x) dy is well-defined for all x ∈ Rn

the integral
Rn
due to the Hölder continuity of f . In particular, it is essential that second-
derivatives of Φ are multiplying the difference [f (y) − f (x)] – the Hölder
continuity of f “cancels” the singular nature of D2 Φ near the origin, at
least enough so that the integral converges.
The presence of the boundary integral on the right-hand side of (4.5) is
necessary in order to cancel the effect of the subtraction of f (x) from f (y).

Proof of Theorem 4.12. To see that (4.6) follows from (4.5), notice that
∆u = u,ii and that according to (4.5),
Z Z

u,ii (x) = Φ,ii (x − y) f (y) − f (x) dy − f (x) Φ,i (x − y)Ni dSy .
Ω0 ∂Ω0

Since Φ,ii (x − y) = 0 if x 6= y , and Φ,ii (x − ·) f (·) − f (x) ∈ L1 (Rn ) , the




first integral on the right-hand side vanishes. Thus


Z
u,ii (x) = −f (x) Φ,i (x − y)Ni dSy . (4.7)
∂Ω0
162 CHAPTER 4. The Laplace/Poisson Equations

Choose R > 0 sufficiently large (for example, R = diam(Ω) + 1) so that


Ω0 ⊂⊂B(x, R), and let A = B(x, R) − Ω0 . Since ∆Φ(x − ·) = 0 in A, the
divergence theorem implies that
Z Z Z
0= ∆Φ(x−y)dx = DΦ(x−y)·n(y)dSy + DΦ(x−y)·n(y)dSy ,
A ∂B(x,R) ∂Ω0

where n denotes the outward-pointing unit normal to ∂A . Since n = −N


on ∂Ω0 , substitution into (4.7) shows that
Z
u,ii (x) = −f (x) Φ,i (x − y)Ni dSy
∂B(x,R)
−(xi − yi ) yi − x i
Z
= −f (x) |x − y|−n dSy
∂B(x,R) ωn−1 R
1
Z
= −f (x) R1−n dSy = −f (x) .
ωn−1 ∂B(x,R)

Next, we establish (4.5). Following the proof of Lemma 4.10, we define


Z  |x − y| 
i
v (x) = ρ Φ,i (x − y)f (y)dy .
Rn 
The derivative of this integrand has an L1 (Rn ) dominating function, so the
dominated convergence theorem allows to differentiate under the integral.
We thus find that
∂  |x − y| i |x − y| 
Z h Z 
vi ,j (x) = ρ Φ,i (x − y)f (y)dy + ρ Φ,ij (x − y)f (y)dy
Ω0 ∂xj  Ω0 
∂  |x − y| i
Z h

= ρ Φ,i (x − y) f (y) − f (x) dy
Ω0 ∂xj 
∂  |x − y| i
Z h
− f (x) ρ Φ,i (x − y)dy
Ω0 ∂yj 
Z  |x − y|  
+ ρ Φ,ij (x − y) f (y) − f (x) dy
Ω0 
Z  |x − y| 
+ f (x) ρ Φ,ij (x − y)dy .
Ω0 

We will show that vi ,j converges to the right-hand side of (4.5) uniformly.

Since Φ,ij (x − y) = − Φ,i (x − y) , integration by parts shows that
∂yj
Z  |x − y| 
ρ Φ,ij (x − y)dy
Ω0 
∂  |x − y| i  |x − y| 
Z h Z
= ρ Φ,i (x − y)dy − ρ Φ,i (x − y)Nj dSy ,
Ω0 ∂yj  ∂Ω0 
4.2. GREEN’S FUNCTION 163

and hence that


Z h ∂  |x − y| i
vi ,j (x) =

ρ Φ,i (x − y) f (y) − f (x) dy
Ω0 ∂xj 
Z  |x − y|  
+ ρ Φ,ij (x − y) f (y) − f (x) dy
Ω0 
Z  |x − y| 
− f (x) ρ Φ,i (x − y)Nj dSy .
∂Ω0 
Following the proof of Lemma 4.10, by the Hölder continuity of f ,
Z h ∂  |x − y| i 
ρ Φ,i (x − y) f (y) − f (x) dy

Ω0 ∂xj 

Z ∂  |x − y| 
≤ ρ Φ,i (x − y) f (y) − f (x) dy

B(x,R) ∂xj 

Z x − y  |x − y| 
i i 0
≤ ρ Φ,i (x − y) f (y) − f (x) dy

<|x−y|<2 |x − y| 

Z 2[f ]C 0,α (Ω)
≤ |x − y|1−n+α dy
<|x−y|<2 ω n−1
2[f ]C 0,α (Ω) Z Z 2
= rα drdSz → 0 as  → 0 .
ωn−1 |z|=1 
Here we note that since R can be chosen independent of x ∈ Ω , the conver-
gence above is in fact uniform in x . Similarly, by (4.1b),
Z  |x − y|  
1 − ρ Φ,ij (x − y) f (y) − f (x) dy

Ω0 
n[f ]C 0,α (Ω) Z Z 2
≤ rα−1 drdSz → 0 as  → 0 ;
ωn−1 |z|=1 0

and again the convergence above is uniform in x ∈ Ω . Consequently, vi ,j


converges to the right-hand side of (4.5) uniformly as  → 0 .
It remains to show that vi ,j → u,ij as  → 0. From Lemma 4.10, vi → u,i
uniformly as  → 0, so using the fundamental theorem of calculus (as in the
proof of Lemma 4.10), we indeed see that u,ij must be equal to the right-
hand side of (4.5). 

4.2 Green’s function

For a point y ∈ Ω, the function Φ(x − y) is harmonic if x 6= y . Therefore,


letting v(x) = Φ(x − y) in Green’s second identity (1.2),
∂u ∂Φ
Z Z Z
Φ(x − y)∆u(x)dx = Φ(x − y) (x)dSx − u(x) (x − y)dSx
Ωk ∂Ωk ∂N ∂Ωk ∂N
164 CHAPTER 4. The Laplace/Poisson Equations

where Ωk = Ω\B(y, k1 ) . For k big enough, ∂Ωk = ∂Ω∪∂B(y, k1 ) . Therefore,


∂Φ ∂u ∂u
Z Z Z
u(x) (x − y)dSx = Φ(x − y) (x)dSx + Φ(x − y) (x)dSx
1
∂B(y, k ) ∂N ∂Ω ∂N 1
∂B(y, k ) ∂N
∂Φ
Z Z
− u(x) (x − y)dSx − Φ(x − y)∆u(x)dx ,
∂Ω ∂N 1
Ω\B(y, k )

here we have to note that the unit normal on ∂B(y, k1 ) points to the center
y. For x ∈ ∂B(y, k1 ) ,

1


 log k if n = 2
 2π
Φ(x − y) = ,
k n−2
if n ≥ 3



(n − 2)ωn−1
∂Φ k n−1
(x − y) = DΦ(x − y) · k(y − x) = .
∂N ωn−1
Therefore, as k → ∞ ,
∂Φ ∂u
Z Z
u(x) (x − y)dSx → u(y) and Φ(x − y) (x)dSx → 0 .
1
∂B(y, k ) ∂N 1
∂B(y, k ) ∂N
Z Z
Moreover, Φ(x − y)∆u(x)dx → Φ(x − y)∆u(x)dx as k → ∞ . As a con-
Ωk Ω
sequence, u ∈ C 2 (Ω) satisfies
∂u ∂Φ
Z Z Z
u(x) = Φ(y − x) (y)dSy − u(y) (y − x)dSy − Φ(y − x)∆u(y)dy .
∂Ω ∂N ∂Ω ∂N Ω
(4.8)
Z
Remark 4.14. The integral Φ(x − y)f (y)dy is called the Newtonian

potential with density f .

Given the formula (4.8) it is tempting to believe that the equation

−∆u = f in Ω,
u=g on ∂Ω ,
∂u
=h on ∂Ω
∂N
has a solution
∂Φ
Z Z Z
u(x) = Φ(x − y)h(y)dSy − g(y) (y − x)dSy + Φ(y − x)f (y)dy .
∂Ω ∂Ω ∂N Ω

This is, in fact, not the case, and we shall examine this in great detail below.
4.3. PROPERTIES OF HARMONIC FUNCTIONS 165

4.3 Properties of harmonic functions


4.3.1 Mean-value property

Now we consider harmonic functions on an open set Ω ⊆ Rn .

Theorem 4.15. If u ∈ C 2 (Ω) is harmonic, then


Z Z
u(x) = − u(y)dSy = − u(y)dy (4.9)
∂B(x,r) B(x,r)

for each ball B(x, r) ⊆ Ω .

Proof. We begin by proving the first equality in (4.9). By the divergence


theorem,
∂u ∂
Z Z
0= (y)dSy = rn−1 u(x + rw)dSw
∂B(x,r) ∂N ∂B(0,1) ∂r
∂ n−1 ∂
Z h Z i
n−1 1−n
=r u(x + rw)dSw = r r u(y)dSy .
∂r ∂B(0,1) ∂r ∂B(x,r)

Therefore,
Z Z
r1−n u(y)dSy = lim r1−n u(y)dSy = ωn−1 u(x) . (4.10)
∂B(x,r) r→0+ ∂B(x,r)

The second equality in (4.9) is obtained by integrating (6.44) in r . 

Exercise: If u ∈ C 2 (Ω) satisfying ∆u ≥ 0 (or ∆u ≤ 0), show that


Z  Z 
u(x) ≤ − u(y)dSy or u(x) ≥ − u(y)dSy ,
∂B(x,r) ∂B(x,r)
Z  Z 
u(x) ≤ − u(y)dy or u(x) ≥ − u(y)dy .
B(x,r) B(x,r)

A function u ∈ C 2 (Ω) is called a sub-harmonic/super-harmonic function if


∆u ≥ 0/≤ 0.

Theorem 4.16 (Converse to mean-value property). If u ∈ C 2 (Ω) satisfies


(4.9) for each B(x, r) ⊆ Ω , then u is harmonic.

Proof. If ∆u 6= 0, there exists some ball B(x, r) ⊂ Ω on which ∆u > 0 (or


R
perhaps ∆u < 0) in B(x, r). But this would imply that 0 = −B(x,r) ∆u(y)dy >
0, a contradiction. 
166 CHAPTER 4. The Laplace/Poisson Equations

Theorem 4.17. If u ∈ L1loc (Ω) satisfies the mean-value property (4.9) for
each B(x, r) ⊆ Ω , then u ∈ C ∞ (Ω) .

Proof. Let η be the standard mollifier defined in Definition 2.25, and let
u = η ∗ u in Ω . Then
Z
1
Z  |x − y| 
u (x) = η (|x − y|)u(y)dy = n η u(y)dy
Ω  B(x,) 
1  r
Z Z
= n η( )u(x + rw)rn−1 dSw dr
 0 ∂B(0,1) 
Z 
1 r
= n ωn−1 u(x) η( )rn−1 dr = u(x) .
 0 
Therefore, u = u in Ω . 

Corollary 4.18. A harmonic function is a C ∞ -function.

Corollary 4.19. The limit of a uniformly convergent sequence of harmonic


functions is harmonic.

4.3.2 Maximum principles

Theorem 4.20 (Strong maximum principle). Suppose that u ∈ C 2 (Ω) ∩


C 0 (Ω) is harmonic within a bounded domain Ω .

(1) Then max u = max u .


Ω ∂Ω

(2) Furthermore, if Ω is connected and there exists a point x0 ∈ Ω such


that u(x0 ) = max u , then u = u(x0 ) within Ω .

Proof. Let M = max u (the boundedness of Ω implies that such an M



exists), and A = {x ∈ Ω | u(x) = M } . The continuity of u implies that A
is closed (relative to Ω). If A is empty, then the maximum of u is attained
on ∂Ω , so we may assume that A is not empty.
Let x ∈ A , and r > 0 be such that B(x, r) ⊆ Ω . According to (4.9)
Z
M = u(x) = − u(y)dy ≤ M .
B(x,r)

Equality can only hold if u(y) = u(x) for all y ∈ B(x, r); thus, B(x, r) ⊆ A,
and hence A is also open in Ω . It must be that A = Ω since it is the only
subset of Ω which is both open and closed in Ω . 
§4.3 Properties of harmonic functions 167

Remark 4.21. In the statement of Theorem 4.20, it suffices to assume that


∆u ≥ 0.

Exercise: Let u be sub-harmonic/super-harmonic in Ω . Then

(1) Then max u = max u/ min u = min u .


Ω ∂Ω Ω ∂Ω

(2) If Ω is connected and there exists a point x0 ∈ Ω such that u(x0 ) =


max u/u(x0 ) = min u , then u = u(x0 ) within Ω .
Ω Ω

Corollary 4.22 (Uniqueness). Let g ∈ C 0 (∂Ω) , f ∈ C 0 (Ω) . Then there


exists at most one solution u ∈ C 2 (Ω)∩C 0 (Ω) to the boundary-value problem

−∆u = f in Ω,
u=g on ∂Ω .

Corollary 4.23 (Comparison). For u, v ∈ C 2 (Ω) ∩ C 0 (Ω), if ∆u ≥ 0,


∆v = 0 in Ω, and u = v on ∂Ω, then u ≤ v in Ω. (For this reason, we call
u subharmonic.

4.3.3 The Harnack inequality

Theorem 4.24. Let u be a non-negative harmonic function in Ω . Then for


any bounded sub-domain Ω0 ⊂⊂Ω , there exists a constant C depending only
on n , Ω0 and Ω such that

max
0
u ≤ C min
0
u.
Ω Ω

Proof. Let y ∈ Ω , B(y, 4R) ⊆ Ω . Then for any two points x1 , x2 ∈ B(y, R) ,
we have
1
Z Z
u(x1 ) = − u(x)dx ≤ u(x)dx ,
B(x1 ,R) |B(0, R)| B(x1 ,2R)
1
Z Z
u(x2 ) = − u(x)dx ≥ u(x)dx .
B(x2 ,3R) |B(0, 3R)| B(x1 ,2R)

Consequently, max u ≤ 3n min u . The general case follows from connect-


B(y,R) B(y,R)
m
ing any two points in Ω0 by an arc Γ so that Γ ⊆
S
B(xk , 4R) ⊆ Ω for
k=1
some xk ∈ Γ and R > 0 . The number m only depends on Ω0 and Ω , so the
constant C depends only on n , Ω0 and Ω . 
168 CHAPTER 4. The Laplace/Poisson Equations

4.3.4 Local estimates

Theorem 4.25. Assume u is harmonic in Ω and U ⊂⊂Ω . Then for any


multi-index α we have
 n|α| |α|
sup |Dα u(x)| ≤ sup |u(x)| , (4.11)
x∈U d x∈Ω

where d = dist(U, ∂Ω) .

Proof. Suppose that α = ei = ( 0, · · · , 0 , 1, 0, · · · , 0) . Since Dα u is har-


| {z }
(i − 1) zeros
monic,
1 1
Z Z
α α
D u(x) = D u(y)dy = div(uei )dy
|B(x, r)| B(x,r) |B(x, r)| B(x,r)
1 |∂B(x, r)|
Z
= u(y)Ni (y)dSy ≤ max |u(y)| .
|B(x, r)| ∂B(x,r) |B(x, r)| y∈B(x,r)
[
As a consequence, with Ur denoting the set B(x, r) ,
x∈U

n
max |Dα u(x)| ≤ max |u(x)| .
x∈U r x∈Ur

The general result follows from applying the above inequality |α| times with
d
r= . 
|α|
Remark 4.26. Inequality (4.11) is also called the gradient estimate for
harmonic functions.

4.3.5 Regularity of weakly harmonic functions

Theorem 4.27. For Ω ⊂ Rn , suppose that u ∈ L1loc (Ω) and satisfies


Z
u(x)∆φ(x)dx = 0 ∀ φ ∈ C02 (Ω) .

Then u is harmonic in Ω.

Proof. Without loss of generality, we may assume that u ∈ L1 (Ω).


Choose  > 0 sufficiently small so that φ := η ∗ φ ∈ C0∞ (Ω), where η
denotes the standard mollifiers given in Definition 2.25. By assumption,
Z Z
0= u(x)∆φ (x)dx = u (x)∆φ(x)dx . (4.12)
Ω Ω
§4.3 Properties of harmonic functions 169

Since u is smooth, we can integrate by parts to find that


Z
∆u (x)φ(x)dx ∀ φ ∈ C02 (Ω) .

Thus, ∆u = 0 in Ω, so u is harmonic.


By Young’s inequality the sequence u is uniformly bounded in L1 (Ω):

ku kL1 (Ω) ≤ kη kL1 (Ω) kukL1 (Ω) = kukL1 (Ω) .

Since u is harmonic,

1
Z

u (x) = u (y)dy ,
ωn Rn B(x,R)

which implies that


1
|u (x)| ≤ ku kL1 (Ω) .
ωn Rn
The local gradient estimate (4.11) then provides the inequality

sup |Dα u(x)| ≤ C · sup |u | ≤ CkukL1 (Ω) .


x∈U Ω

We have therefore shown that Dα u is bounded and equicontinuous on


any U ⊂⊂Ω. By the Arzela-Ascoli Compactness Theorem 2.110, there exists
0
some subsequence 0 such that u → v ∈ C 2 (U ).
0
On the other hand, u → u in L1 (U ) so u = v on U . By pushing ∂U
closer to ∂Ω, we see that u is smooth in Ω. Hence for all φ ∈ C02 (Ω),
Z Z
0= u(x)∆φ(x)dx = ∆u(x)φ(x)dx ,
Ω Ω

so that u is harmonic. 

4.3.6 Liouville’s Theorem

Theorem 4.28. Suppose u : Rn → R is harmonic and bounded. Then u is


constant.

Proof. By gradient estimate (4.11), the derivatives Dα u(x) have to vanish


for all x ∈ Rn and multi-index α . In particular, this implies that u is
constant along any lines, so u is constant. 
170 CHAPTER 4. The Laplace/Poisson Equations

Corollary 4.29. Let n ≥ 3 and f ∈ C 0,α (Rn ) with compact support in Rn .


Then any bounded solution of

−∆u = f in Rn

has the form


Z
u(x) = Φ(x − y)f (y)dy + C
Rn

for some constant C .

4.3.7 Analyticity

Theorem 4.30. Assume u is harmonic in Ω . Then u is analytic in Ω ,


that is, for each x0 ∈ Ω ,
X Dα u(x0 )
u(x) = (x − x0 )α
α
α!

within some ball B(x0 , r) .

4.4 Computing Green’s functions

For fixed x ∈ Ω , suppose that there is a harmonic function Φ̃x (y) such that
∆y Φ̃x (y) = 0 for all y ∈ Ω and

Φ̃x (y) = Φ(y − x) ∀ y ∈ ∂Ω .

Then by Green’s second identity,


∂u ∂ Φ̃x i
Z Z h
Φ̃x (y)∆u(y)dy = Φ̃x (y) (y) − u(y) (y) dSy . (4.13)
Ω ∂Ω ∂N ∂N
By (4.8) and (4.13), we obtain that if u ∈ C 2 (Ω) ∩ C 0 (Ω) , then
∂G
Z Z
u(x) = − u(y) (x − y)dSy − G(x, y)∆u(y)dy , (4.14)
∂Ω ∂N Ω

where G(x, y) = Φ(y −x)− Φ̃x (y). The function G(x, y) is called the Green’s
function for the domain Ω.

Theorem 4.31 (Symmetry of Green’s function). For all x , y ∈ Ω , x 6= y ,


we have

G(x, y) = G(y, x) .
§4.4 Computing Green’s functions 171

Proof. Fix x, y ∈ Ω , x 6= y . Define v(z) = G(x, z) and w(z) = G(y, z) . The


goal is to show that v(y) = w(x) .
Letting Ω = Ω−(B(x, )∪B(y, )), and applying Green’s second identity
(1.2), we obtain that
Z h ∂v ∂w i
Z h ∂w ∂v i
w− v dSz = v− w dS ,
∂B(x,) ∂N ∂N ∂B(y,) ∂N ∂N

where N denotes the inward-pointing unit normal on ∂B(x, ) ∪ ∂B(y, ) .


Passing  → 0 , the left-hand side converges to w(x) while the right-hand
side converges to v(y) . 

4.4.1 The case Ω = Rn−1 × R+

For x ∈ Rn+ ≡ Rn−1 × R+ , let x̃ = (x1 , · · · , xn−1 , −xn ) . Then Φ̃x (y) =
Φ(y − x̃) is harmonic in Rn+ , and Φ̃x (y) = Φ(y − x) for all y ∈ ∂Rn+ .
Therefore, the Green’s function is given by

G(x, y) = Φ(y − x) − Φ(y − x̃) ,

the outward unit normal to ∂Rn+ is N = −en = (0, ..., 1) so ∂G/∂N =


−∂G/∂yn , and
∂G −1 2xn
(x, y) = ,

∂N yn =0 ωn−1 |x − y|n

and Green’s representation formula (4.14) suggests that the (bounded) so-
lution to

∆u = 0 in Rn+ , (4.15a)
u=f on ∂Rn+ (4.15b)

is

∂G 2xn f (y)
Z Z
u(x) = − f (y) (x, y)dSy = dy (∀ x ∈ Rn+ ) . (4.16)
n
∂R+ ∂N ωn−1 Rn−1 |x − y|n

2xn
With K(x, y) = , the equation (4.16) can be written as
ωn−1 |x − y|n
Z
u(x) = K(x, y)f (y)dSy .
∂Rn
+
172 CHAPTER 4. The Laplace/Poisson Equations

The function K is termed the Poisson kernel for Rn+ , and (4.16) is called
the Poisson integral formula.
It remains to verify that (4.16) provides a solution u ∈ C 2 (Rn+ ) ∩ C 0 (Rn+ )
with prescribed boundary condition.

Theorem 4.32. Assume that f ∈ C 0 (∂Rn+ ) ∩ L∞ (∂Rn+ ) , and u is given by


(4.16). Then u ∈ C 2 (Rn+ ) ∩ C 0 (Rn+ ) satisfies (4.15a) and (4.15b).

Proof. Since K(x, y) is smooth if x 6= y , (4.16) indeed shows that u is


smooth . In particular, u ∈ C 2 (Rn+ ) .
Next, we note that the Poisson kernel is normalized; namely,
2xn 1
Z Z
K(x, y)dSy = n
dy = 1 ∀ x ∈ Rn+ . (4.17)
n
∂R+ ωn−1 R n−1 |x − y|

The identity (4.17) is an immediate consequence of (4.14) with u ∈ C 2 (Rn+ )∩


C 0 (Rn+ ) taken to be u(x) = 1.
Let z ∈ ∂Rn+ . Since f ∈ C 0 (∂Rn+ ) , given  > 0 , there exists δ > 0 such
that

|f (x) − f (z)| < whenever |x − z| < 2δ .
2
Then if |x − z| < δ , x ∈ Rn+ ,
Z 
|u(x) − f (z)| = K(x, y) f (y) − f (z) dSy

∂Rn
+
Z Z
≤ K(x, y)|f (y) − f (z)|dSy + 2 max
n
|f | K(x, y)dSy
∂Rn ∂R+ ∂Rn
+ ∩B(z,2δ) + \B(z,2δ)


Z
≤ + 2 max |f | K(x, y)dSy .
2 ∂Rn
+ ∂Rn
+ \B(z,2δ)
| {z }
≡I

If y ∈ ∂Rn+ \B(z, 2δ) , |y − z| > 2δ . So if |x − z| < δ1 ≤ δ ,


1
|y − z| ≤ |y − x| + |x − z| ≤ |y − x| + |y − z|
2
which implies |y − x| ≥ 12 |y − z| . As a consequence,
2n+2 xn Cxn
Z
I≤ max |f | n
dSy ≤
ωn−1 ∂R+ n
∂Rn
+ \B(z,2δ)
|y − z| δ

for some constant C < ∞ . Choose δ1 even smaller so that 2Cδ1 < δ , then
|u(x) − f (z)| <  whenever |x − z| < δ1 . This proves u ∈ C 0 (Ω) . 
§4.4 Computing Green’s functions 173

4.4.2 The case Ω = B(0, 1) or Ω = B(0, R)


x
For x ∈ B(0, 1) , let x̃ = , and Φ̃x (y) = Φ(|x|(y − x̃)) . Then Φ̃x (y)
|x|2
is harmonic in Ω , i.e., ∆y Φ̃x (y) = 0 for all x, y ∈ Ω . Moreover, Φ̃x (y) =
Φ(y − x) for all y ∈ ∂B(0, 1) . Therefore, the Green’s function for the unit
ball is

G(x, y) = Φ(x − y) − Φ(|x|(y − x̃)) .

Using (4.14), we find that the solution to

∆u = 0 in B(0, 1) ,
u=f on ∂B(0, 1)

is
∂G 1 − |x|2 f (y)
Z Z
u(x) = − f (y) (x, y)dSy = dSy .
∂B(0,1) ∂N ωn−1 ∂B(0,1) |x − y|n

since
∂G −1 1 − |x|2
(x, y) = .

∂N y∈∂B(0,1) ωn−1 |x − y|n
By a change of variables, the solution to

∆u = 0 in B(0, R) , (4.18a)
u=f on ∂B(0, R) , (4.18b)

is then
R2 − |x|2 f (y)
Z
u(x) = dSy . (4.19)
ωn−1 R ∂B(0,R) |x − y|n

Similar to Theorem 4.32, we have that

Theorem 4.33. Assume that f ∈ C 0 (∂B(0, R)) , and u is given by (4.19).


Then u ∈ C 2 (B(0, R)) ∩ C 0 (B(0, R)) satisfies (4.18a) and (4.18b).

The function
R2 − |x|2 1
K(x, y) =
ωn−1 R |x − y|n
is the Poisson kernel for the ball B(0, R) .
174 CHAPTER 4. The Laplace/Poisson Equations

4.5 Perron’s method and solutions to the Poisson


problem1

In this section, we prove the existence of solutions to

−∆u = f in Ω, (4.20a)
u=g on ∂Ω . (4.20b)

using Perron’s method under the assumption that f ∈ C 0,α (Ω) and g ∈
C 0 (∂Ω) .
First, we extend f to whole Rn with compact support so that the exten-
sion, still denoted by f , belongs to C 0,α (Rn ) . Let ϕ = Φ∗f , and v = u−ϕ .
By Lemma 4.12, −∆ϕ = f , v is harmonic. So, v solves

−∆v = 0 in Ω, (4.21a)
v =g−ϕ≡ψ on ∂Ω . (4.21b)

As long as we know how to solve the Dirichlet problem (4.21), we obtain a


solution to (4.20) by summing ϕ and the solution to (4.21). Therefore, we
concentrate on how (4.21) is solved.
First we generalize the notion of sub-harmonic function. Recall that a
function w ∈ C 2 (Ω) is sub-harmonic if ∆w ≥ 0 in Ω , and
Z
w(ξ) ≤ − w(y)dSy ≡ Mw (ξ, ρ) . (4.22)
∂B(ξ,ρ)

A function w ∈ C 0 (Ω) is called sub-harmonic, if for each ξ ∈ Ω , (4.22)


holds for all ρ > 0 such that B(ξ, ρ)⊂⊂Ω . Let σ(Ω) denote the space of all
sub-harmonic functions on Ω , and given ψ ∈ C 0 (∂Ω) , define
n o
σψ (Ω) = w ∈ σ(Ω) ∩ C 0 (Ω) w ≤ ψ on ∂Ω .

σψ (Ω) is non-empty since the constant function w ≡ c belongs to σψ (Ω) if


c ≤ inf ψ(x) .
x∈∂Ω
For u ∈ C 0 (Ω) and B(ξ, ρ)⊂⊂Ω , we define uξ,ρ , the harmonic lifting of
u, as the function in C 0 (Ω) for which

∆x uξ,ρ (x) = 0 in B(ξ, ρ) ,


uξ,ρ (x) = u(x) in Ω\B(ξ, ρ) .
1
The reader may skip this section on the first reading
§4.5 Perron’s method and solutions to the Poisson problem 175

Claim: For u ∈ σ(Ω) and B(ξ, ρ)⊂⊂Ω , u(x) ≤ uξ,ρ (x) for all x ∈ Ω , and
uξ,ρ ∈ σ(Ω) .

Proof. It suffices to show that

uξ,ρ (x) ≤ Muξ,ρ (x, r) ∀ r > 0 such that B(x, r)⊂⊂Ω . (4.23)

If B(x, r) ⊆ B(ξ, ρ) , since uξ,ρ is harmonic in B(ξ, ρ) , (4.23) holds because of


the mean-value property for the harmonic functions. If B(x, r)∩B(ξ, ρ) = ∅ ,
then uξ,ρ = u , so (4.23) holds because of (4.22). Other than these two
cases, let w be the harmonic function satisfying w = uξ,ρ on ∂B(x, r) . On
∂B(x, r) ∩ B(ξ, ρ) , w = uξ,ρ ≥ u , and on ∂B(x, r)\B(ξ, ρ) , w = uξ,ρ = u ,
w ≥ u on ∂B(x, r) , which by the maximum principle implies that w ≥ u in
B(x, r) . Apply the maximum principle once again to the domain B(x, r) ∩
B(ξ, ρ) and B(x, r)\B(ξ, ρ) , we conclude that w ≥ uξ,ρ in B(x, r) and hence
(4.23) holds. 

Claim: Let u1 , · · · , uk ∈ σψ (Ω) , and v = max{u1 , · · · , uk } . Then v ∈


σψ (Ω) .

Proof. Given ξ ∈ Ω , for all sufficient small ρ such that B(ξ, ρ)⊂⊂Ω ,

v(ξ) = max{u1 (ξ), · · · , uk (ξ)} ≤ max{Mu1 (ξ, ρ), · · · , Muk (ξ, ρ)} ≤ Mv (ξ, ρ) .


Claim: For all given ψ ∈ C 0 (∂Ω) , the function wψ (x) ≡ sup w(x) is
w∈σψ (Ω)
well-defined and is harmonic in Ω .

Proof. wψ defined above is well-defined due to the maximum principle for


sub-harmonic functions, and is clearly in σψ (Ω) since
Z Z
sup − w(y)dSy ≤ − sup w(y)dSy .
w∈σψ (Ω) ∂B(ξ,ρ) ∂B(ξ,ρ) w∈σψ (Ω)

It suffices to show that wψ has the mean-value property. Suppose the con-
trary, then
Z
wψ (ξ) < − wψ (y)dSy
∂B(ξ,ρ)
176 CHAPTER 4. The Laplace/Poisson Equations

for some ξ ∈ Ω and some ρ such that B(ξ, ρ)⊂⊂Ω . By the previous two
claims, the function (wψ )ξ,ρ belongs to σψ (Ω) , and wψ (ξ) < (wψ )ξ,ρ (ξ) .
Then wψ (ξ) 6= sup w(ξ) . 
w∈σψ (Ω)

If (4.21) has a solution, it has to be equal to wψ defined above since the


solution itself belongs to σψ (Ω) . In other words, wψ is the only candidate
for the solution. In order to make sure that wψ solves (4.21), we need to
make sure that wψ satisfies the boundary condition.

Definition 4.34 (Barrier Property). A domain is said to have the barrier


property if for each η ∈ ∂Ω , there exists a function, called a barrier function,
Qη ∈ σ(Ω) ∩ C(Ω) for which

Qη (η) = 0 , Qη (x) < 0 for x ∈ ∂Ω , x 6= η .

The barrier property can be verified for a large class of domains Ω . For
example, if Ω is strictly convex in the sense that through each point η ∈ ∂Ω
there passes a hyperplane πη having only η in common with Ω , then Ω has
the barrier property.
As long as Ω has the barrier property, for each y ∈ ∂Ω , there exists a
sub-harmonic function w ∈ σψ (Ω) and w(y) = ψ(y) (for example, consider
w(x) = ψ(y) + Qy (x)). The only thing it remains to be proved is that
wψ ∈ C 0 (Ω) , or

lim wψ (x) = ψ(η) .


x∈Ω
x→η

Claim: If Ω has the barrier property, then for η ∈ ∂Ω ,

lim inf wψ (x) ≥ ψ(η) .


x∈Ω
x→η

Proof. For  > 0 and K > 0 , the function v(x) = ψ(η) −  + KQη (x) belongs
to σ(Ω) ∩ C 0 (Ω) , and satisfies

v(x) ≤ ψ(η) −  ∀ x ∈ ∂Ω , v(η) = ψ(η) −  .

Since ψ ∈ C 0 (∂Ω) , there exists δ > 0 such that |ψ(x) − ψ(η)| <  whenever
|x − η| < δ , x ∈ ∂Ω ; thus v(x) ≤ ψ(x) if |x − η| < δ . If |x − η| ≥ δ , we
§4.5 Perron’s method and solutions to the Poisson problem 177

can choose K large enough so that v(x) ≤ ψ(x) since Qη has negative upper
bound on |x − η| ≥ δ . Therefore, v ∈ σψ (Ω) (if K is large enough). By the
definition of wψ , v(x) ≤ wψ (x) for all x ∈ Ω ; hence

ψ(η) −  = lim inf v(x) ≤ lim inf wψ (η) . (4.24)


x∈Ω x∈Ω
x→η x→η

We then conclude the claim since (4.24) holds for all  > 0 and all η ∈ ∂Ω .


Claim: If Ω has the barrier property, then for η ∈ ∂Ω ,

lim wψ (x) = ψ(η) .


x∈Ω
x→η

Proof. It suffices to show that

lim sup wψ (x) ≤ ψ(η) . (4.25)


x∈Ω
x→η

This is done by considering −w−ψ (x) which is defined in Ω by

−w−ψ (x) = − sup w(x) = inf v(x) .


w∈σ−ψ (Ω) −v∈σ−ψ (Ω)

For all w ∈ σψ (Ω) and −v ∈ σ−ψ (Ω) , w ≤ ψ ≤ v on ∂Ω , and w − v ∈


σ(Ω) ∩ C 0 (Ω) ; therefore by the maximum principle for the sub-harmonic
functions,

w≤v ∀x ∈ Ω ⇒ wψ (x) ≤ −w−ψ (x) ∀ x ∈ Ω .

By previous claim,
lim inf w−ψ (x) ≥ −ψ(η) ⇒ lim sup wψ (x) ≤ lim sup −w−ψ (x) ≤ ψ(η) . 
x∈Ω x∈Ω x∈Ω
x→η x→η x→η

Theorem 4.35. If the domain Ω has the barrier property, and f ∈ C 0,α (Ω) ,
then there exists a unique solution u ∈ C 2 (Ω) ∩ C 0 (Ω) of the Dirichlet prob-
lem

−∆u = f in Ω
u=g on ∂Ω ,

for arbitrary continuous boundary value g .


178 CHAPTER 4. The Laplace/Poisson Equations

4.6 Exercise

Problem 4.1. Let Ω = B(0, 12 ) ⊆ R2 denote the open ball of radius 1


2
centered at the origin. For x = (x1 , x2 ) ∈ Ω , let
" ! #
1
u(x1 , x2 ) = x1 x2 log log p 2 − log log 2 .
x1 + x22

(a) Show that u ∈ C 1 (Ω) ;

(b) Show that ∆u ∈ C(Ω) for j = 1, 2 , but that u 6∈ C 2 (Ω) .


2 (D) ∩ C(D) be a solution to the problem
Problem 4.2. Let u ∈ Cloc

−∆u = 1 in D ≡ (−1, 1) × (−1, 1) ⊆ R2 ,


u=0 on ∂D .

Show that u cannot belong to C 2 (D) .

Problem 4.3. Find a solution to the Dirichlet problem ∆u = 0 in the square

(x, y) ∈ R2 − 1 ≤ x ≤ 1 , −1 ≤ y ≤ 1


satisfying the boundary conditions


 3π 
u(x, y) = cos x , on y = ±1 , −1 ≤ x ≤ 1 ,
2
 3π 
u(x, y) = cos y , on x = ±1 , −1 ≤ y ≤ 1 .
2
Problem 4.4. Let u , v be smooth harmonic functions, such that

u(tx) ≡ ta u(x) , v(tx) = tb v(x)

for all x ∈ Rn , t > 0 , with constants a 6= b . Use Green’s identity to show


that
Z
uvdS = 0 .
∂B(0,1)

Problem 4.5. Let u be a harmonic function in the unit ball B1 ≡ B(x0 , 1) ⊂


Rn . Prove the following gradient estimate:
h i
|Du(x0 )| ≤ n sup u(x) − u(x0 )
x∈B1
§4.6 Exercise 179

Hint: Note that all the derivatives ∂xi u are harmonic in B1 , so that by the
mean value and divergence theorems,
∂u 1 ∂u 1
Z Z
(x0 ) = (x)dx = uNi dS ,
∂xi |B1 | B1 ∂xi |B1 | ∂B1
where Ni is the i-th component of the unit normal N to ∂B1 . Obviously, for
x ∈ ∂B1 , we have N(x) = x − x0 .
2 (R2 ) ∩ C (R2 )
Problem 4.6. Show that there are no functions u ∈ Cloc + loc +
satisfying the properties

u ≥ 0, ∆u = 0 in R2+ ≡ {x = (x1 , x2 ) ∈ R2 | x2 > 0} , u(x1 , 0) = x21 .

Hint: Suppose there exists such a function u . For arbitrary R > 0 , compare
2 (R2 ) ∩ C(R2 ) to the problem
u with the solution v ∈ Cloc + +

∆v = 0 in R2+ , v(x1 , 0) = ζ(x1 )x21 ,

where

ζ ∈ Cc∞ (−2R, 2R) , 0 ≤ ζ ≤ 1, and ζ ≡ 1 on [−R, R] .

Problem 4.7. Let u(x1 , x2 ) ∈ C 2 (Ω) , where Ω ≡ {x = (x1 , x2 ) ∈ R2 | x1 >


0 , x2 > 0} , such that

∆u = 0 in Ω, u=0 on ∂Ω .

In addition, let |u(x)| ≤ c1 +c2 |x| in Ω with some constants c1 and c2 . Show
that u ≡ 0 in Ω .
Hint: Extend the domain where u is defined to R2 and use the gradient
estimates to show that u is in fact a linear function.

Problem 4.8. Let u(x) = u(x1 , x2 ) be a bounded solution of the Laplace


equation

∆u = 0 in R2+

with boundary condition


|x1 |
u(x1 , 0) = g(x1 ) = .
1 + x21
Show that the gradient Du is unbounded on R2+ .
180 CHAPTER 4. The Laplace/Poisson Equations

Problem 4.9. Find the Green function for the domain Ω = x ∈ Rn |x|2 <


1 , xn > 0 and the corresponding Green’s representation formula for the
solution of the Dirichlet problem

∆u = 0 in Ω,
u=f on ∂Ω .
Chapter 5

Second Order Linear Elliptic


Equations

5.1 The strong form of 2nd-order elliptic bound-


ary value problems

Let Ω ⊆ Rn denote an open and smooth domain, and suppose that u ∈


C 2 (Ω). We consider the following second-order differential operators defined
in Ω:

∂  ij ∂u  ∂u
(Lu)(x) = − a (x) (x) + bi (x) (x) + c(x)u(x)
∂xi ∂xj ∂xi

= − div a(x) ·Du(x) + b(x) ·Du(x) + c(x)u(x) (divergence form)

or

∂2u ∂u
(Lu)(x) = −aij (x) (x) + bi (x) (x) + c(x)u(x)
∂xi ∂xj ∂xi
= − a(x) :D2 u(x) + b(x) ·Du(x) + c(x)u(x) (non-divergence form).

We remind the reader, that we are employing the Einstein summation con-
vention, in which repeated Latin indices are summed from 1 to n, and re-
peated Greek indices are summed from 1 to n − 1. For example, fi gi =
Pn n−1
P
fi gi and fα gα = fα gα . Also, our intrinsic notation, is to use dots to
i=1 i=1
indicate contraction between indices, so that the scalar inner-product of two
vectors X and Y is denoted by X · Y = Xi Yi , and the trace of the product
of two matrices A and B is written as A : B = Aij Bji .

181
182 CHAPTER 5. Second Order Linear Elliptic Equations

Definition 5.1. The operator L is said to be elliptic at a point x ∈ Ω if


there exist positive functions λ(x) and Λ(x) such that

0 < λ(x)|ξ|2 ≤ aij (x)ξj ξi ≤ Λ(x)|ξ|2 (5.1)

for all ξ = (ξ1 , · · · , ξn ) ∈ Rn − {0}.


If there exists positive constants λ 0 and Λ0 such that λ(x) ≥ λ 0 > 0 and
Λ(x) ≤ Λ0 < ∞, then L is uniformly elliptic in Ω.

Definition 5.2 (Elliptic Boundary Value Problem). Given an (uniformly)


elliptic operator L, we say that u : Ω → R is a strong solution of an elliptic
boundary value problem, or BVP, if

(a) Lu(x) = f (x) for almost all x ∈ Ω;

(b) B∂Ω u(x) = g(x) for almost all x ∈ ∂Ω,

where B∂Ω denotes a “boundary” operator, f is the interior forcing function,


and g is the prescribed boundary condition.

While it is possible to impose a wide variety of boundary conditions,


perhaps the most fundamental are known as the Dirichlet, Neumann, and
periodic boundary conditions. Here is a partial list of boundary conditions
that we shall examine:

1. B∂Ω u = u on ∂Ω (Dirichlet boundary condition);


∂u
2. B∂Ω u = aij Ni on ∂Ω (Neumann boundary condition);
∂xj
3. with ∂Ω = Γ1 ∪ Γ2 so that Γ1 ∩ Γ2 = ∅, we set
∂u
B∂Ω u = 1Γ1 u + 1Γ2 aij Ni (Mixed boundary conditions);
∂xj
∂u
4. B∂Ω u = aij Ni + ku on ∂Ω (Robin boundary condition);
∂xj
5. with g(x) = 0 and Ω = Tn , and with ei denoting the n orthonormal
unit vectors,
B∂Ω u(x) = u(x + ei ) − u(x) (Periodic boundary conditions);

We will additionally describe certain non standard boundary conditions,


involving elliptic differential operators on the boundary ∂Ω.
5.1. STRONG SOLUTIONS 183

5.1.1 The Dirichlet problem

Given an interior forcing function f : Ω → R and a function g : ∂Ω → R


on the boundary, the solution to the Dirichlet problem on Ω is a function
u : Ω → R which satisfies

Lu = f in Ω, (5.2a)
u=g on ∂Ω . (5.2b)

The Dirichlet boundary condition is also known as an essential boundary


condition. This is due to the inclusion of the boundary condition into the
functional framework used for existence theorems (as will be made precise
below).

5.1.2 The Neumann problem

Given an interior forcing function f : Ω → R and a function g : ∂Ω → R


on the boundary, the solution to the Neumann problem on Ω is a function
u : Ω → R which satisfies

Lu = f in Ω, (5.3a)
∂u
aij Ni = g on ∂Ω . (5.3b)
∂xj

The Neumann boundary condition is also known as a natural boundary con-


dition. Unlike the Dirichlet boundary condition, this boundary condition
appears naturally in the variational form of the boundary value problem.

5.1.3 The Robin problem

It is also to possible to impose certain linear combinations of the Dirichlet


and Neumann boundary conditions. In this case, given an interior forcing
function f : Ω → R and a function g : ∂Ω → R on the boundary, the solution
to the Robin problem on Ω is a function u : Ω → R which satisfies

Lu = f in Ω, (5.4a)
∂u
aij Ni + ku = g on ∂Ω . (5.4b)
∂xj

The sign of k will play a crucial role in existence theorems.


184 CHAPTER 5. Second Order Linear Elliptic Equations

5.1.4 A non-standard elliptic problem

Motivated by problems involving surface tension (and more generally, surfac-


tant theory), we shall also analyze a non standard boundary value problem,
in which a Neumann-like boundary operator is supplemented by a “surface”
operator (to be read as boundary operator) which is itself elliptic.
A prototype BVP is the following: Given an interior forcing function
f : Ω → R and a function g : ∂Ω → R on the boundary, find a function
u : Ω → R which satisfies

Lu = f in Ω, (5.5a)
∂u
aij Ni = ∆g u + g on ∂Ω . (5.5b)
∂xj

In this setting, the smooth boundary ∂Ω is to be thought of as (n − 1)-


dimensional Riemannian manifold1 with Riemannian metric g: (∂Ω, g). The
Riemannian metric is a positive definite (n − 1) × (n − 1) matrix which in a
local coordinate basis is denoted by gαβ , where the indices α, β = 1, ..., n−1.
The inverse of the metric g −1 is denoted with raised superscripts g αβ .
The operator ∆g is called the Laplace-Beltrami operator on ∂Ω is defined
as
1 ∂ p ∂u 
∆g u = p det(g)g αβ . (5.6)
det(g) ∂xα ∂xβ

The Laplace-Beltrami operator (5.6) is the natural generalization of the


Laplacian on Rn−1 .

5.1.5 A vector-valued Dirichlet problem: Linear Elasticity

Our examples, thus far, have all been for scalar-valued solutions, but the
majority of problems coming from physics and engineering are indeed vector-
valued. Linear elasticity affords a nice example of a second-order boundary
value problem for a vector valued unknown. The so-called displacement
1
When Ω is a connected and simply connected bounded subset of Rn , then for the case
that n = 2, the boundary ∂Ω is a closed curve and in the case that n = 3, the boundary ∂Ω
is a closed surface. For the former, the reader may suppose that ∂Ω is diffeomorphic to the
circle S1 , while for the latter, the boundary ∂Ω can be thought of as diffeomorphic to the
unit sphere S2 . While both S1 and S2 are particular examples of Riemannian manifolds,
standard polar coordinates can be employed to parameterize such boundaries.
5.1. STRONG SOLUTIONS 185

vector u : Ω ⊆ R3 → R3 , measures the displacement of an elastic material


in the three physical space dimensions. In particular,

u(x) = (u1 (x), u2 (x), u3 (x)) ,

where u1 and u2 capture the amount of in-plane (horizontal) deformation,


and the u3 represents the vertical (or out of plane) deformation of the ma-
terial.
In order to describe the associated BVP problem, we must generalize
the setting of the operator L from possessing a coefficient matrix Aij (which
depends on two indices) to a coefficient tensor C ijk` (which depends on four
indices). In the setting of elasticity, such a tensor encodes the material
response to tension, compression, shear forcing, moments, etc., in all of the
three space directions. In the parlance of continuum mechanics, C encodes
Hooke’s Law relating the strain of an elastic material to the resulting stress.

Definition 5.3 (The fourth-rank tensor C). The fourth-rank tensor C, writ-
ten as C ijk` with respect to the standard coordinate basis in R3 , satisfies the
following properties:

1. Positivity: ∃λ > 0, C ijk` ξi ξj ηk η` ≥ λ|ξ|2 |η|2 for all ξ, η ∈ R3 .

2. Symmetry: C ijk` = C jik` = C ij`k = C k`ij for all i, j, k, ` = 1, 2, 3.

Definition 5.4 (Strain and stress). The strain tensor  is defined to by the
symmetric part of the 3 × 3-matrix Du. In components,
∂uk ∂u`
kl = + .
∂x` ∂xk
The stress tensor σ is then defined as

σ ij = C ijk` k` .
∂uk
Due to the symmetry assumption on C ijk` , we see that σ ij = C ijk` .
∂x`
Definition 5.5 (Divergence of a matrix). Suppose that M ij (x) denotes a
differentiable matrix function on Ω. The divergence of M , is the n-vector
given by
∂M ij
[Div M ]i = .
∂xj
186 CHAPTER 5. Second Order Linear Elliptic Equations

The equations of linear elasticity are the static conservation law

Div σ = f in Ω

for a vector forcing function f : Ω → R3 . More explicitly, we have that

∂  ijk` ∂u` 
− C = fi in Ω, (5.7a)
∂xj ∂xk
(B∂Ω u)i = g i on ∂Ω , (5.7b)

where f i is the interior forcing vector and g i denotes the vector of boundary
data. The operator B can be chosen to represent either the vector Dirichlet,
Neumann, mixed-type boundary conditions, or periodic boundary conditions
can be imposed if the domain Ω = T3 .
Of particular interest to us will be the specification of Dirichlet and
Neumann boundary conditions:

1. [B∂Ω u]i = ui on ∂Ω (Dirichlet boundary condition);

∂uk
2. [B∂Ω u]i = C ijk` Nj on ∂Ω (Neumann boundary condition).
∂x`
Note that prescribing the Dirichlet conditions on ∂Ω requires us to prescribe
the displacement field ui of the elastic material on the boundary ∂Ω. On the
other hand, the Neumann boundary condition, also known as the traction
boundary condition, requires us to instead specify, the so-called traction
vector, defined to be the normal component of the stress tensor, σ ij Nj , on
the boundary ∂Ω.

5.2 Variational Formulations of Elliptic BVPs

The strong formulation of second-order elliptic BVPs as given by Definition


5.2 requires that linear combinations of partial derivatives of the solution
should equal a given forcing function at almost every point. Suppose that
the solution u(x) represents the equilibrium state of the temperature dis-
tribution. Is it possible to measure the temperature at every point of the
continuum, and moreover, is it possible to produce a change in the tempera-
ture (via forced heating) at every point? If measurements of the temperature
are made with a sensor of some sort, then there exists a smallest physical
§5.2 Variational formulations of elliptic BVPs 187

scale, below which it is impossible to discern temperature differences. In par-


ticular, the sensor “averages” about a small region A (with Zdiameter below
the smallest physical scale), and in fact produces the result u(x)φ(x) dx,
A
where φ(x) depends on the type of sensor that is employed.
This motivates to consider weighted averages of elliptic BVPs, and per-
haps the simplest case to consider is the Dirichlet BVP with homogeneous
boundary condition u = 0. To this end, suppose that φ ∈ C0∞ (Ω), and
consider instead of Lu = f in Ω and u = 0 on ∂Ω, the following integrated
equality:
Z Z
Lu(x)φ(x) dx = f (x)φ(x) dx ∀ φ ∈ C0∞ (Ω) . (5.8)
Ω Ω

Clearly, if (5.8) holds, then Lu(x) = f (x) for a.e. x ∈ Ω. On the other hand,
the introduction of the integral over the domain Ω allows us to employ one
of the most powerful tools in analysis: integration by parts. We can rewrite
(5.8) as
Z Z Z
ij i
a (x)u,j (x)φ,i(x) dx + b (x)u,i(x)φ(x) dx + c(x)u(x)φ(x) dx
Ω Z Ω Z Ω

− aij (x)u,j (x)Ni (x)φ(x)dSx = f (x)φ(x) dx ∀ φ ∈ C0∞ (Ω) ,


∂Ω Ω
(5.9)

and as the test function φ vanishes on ∂Ω, the boundary integral in (5.9) is
equal to zero, so that
Z Z Z
ij i
a (x)u,j (x)φ,i(x) dx + b (x)u,i(x)φ(x) dx + c(x)u(x)φ(x) dx
Ω Z Ω Ω

= f (x)φ(x) dx ∀ φ ∈ C0∞ (Ω) . (5.10)


As the coefficient matrix aij ∈ L∞ (Ω), and as our test function φ has square-
integrable partial derivatives, we see that the integral equality (5.10) is well
defined whenever both u and u,j are square-integrable, i.e., u ∈ H 1 (Ω).
Thus, we have reduced the regularity requirements of the solution u,
from C 2 (Ω) to H 1 (Ω). As we will explain below, the integral equality given
in (5.10) (with certain functional constraints) is known as the variational
formulation of the Dirichlet problem. In order for solutions of such a vari-
ational problem to satisfy the homogeneous Dirichlet boundary condition
188 CHAPTER 5. Second Order Linear Elliptic Equations

u = 0, we insist that u be element of a subspace of H 1 (Ω); namely, we


require solutions u of (5.10) to be in the subspace H01 (Ω). Furthermore, we
as C0∞ (Ω) is dense in H01 (Ω), we may consider all test functions to be in
H01 (Ω) as well.
With H01 (Ω) acting as both the space of solutions and the space of test
functions, we have converted a second-order elliptic BVP to a variational
problem in a Hilbert space, a setting very much amenable to the Lax-
Milgram theorem for the existence and uniqueness of solutions. In turn,
the Lax-Milgram theorem is a generalization of the Riesz Representation
theorem.
To explain the choice of the spaces V used for the various boundary
conditions discussed above, it will be convenient to first examine a simpli-
fied version of the operator L in which the lower-order terms vanish. In
particular, we shall assume2 that
 
∂ ij ∂
L=− a (x) .
∂xi ∂xj
Definition 5.6 (Bilinear form associated to L). We let B : V × V → R be
given by
∂u ∂v
Z
B(u, v) = aij (x) dx u, v ∈ V ,
Ω ∂xj ∂xi
where V denotes a subspace of H 1 (Ω).

Definition 5.7 (Weak Solutions to elliptic BVP). Given f ∈ L2 (Ω), say


that u ∈ V is a weak solution of Lu = f in Ω with B∂Ω u = 0 on ∂Ω if
Z
B(u, v) = f (x)v(x) dx ∀ v ∈ V . (5.11)

The choice of subspace V will depend on the type of homogeneous boundary
condition that is imposed. For certain boundary conditions, (5.11) will be
supplemented by an additional bilinear form b : v × v → R, where v denotes
a subspace of H 1 (∂Ω) and will be specified below. If we allow the possibility
that b = 0, then a more general form of (5.11) can be written as follows:
Find u ∈ V ∩ v such that
Z
B(u, v) + b(u, v) = f (x)v(x) dx ∀ v ∈ V ∩ v . (5.12)

2
We will return to the general operator L below.
§5.2 Variational formulations of elliptic BVPs 189

5.2.1 The Dirichlet problem

We begin the discussion of the Dirichlet problem (5.2) with homogeneous


boundary conditions u = 0. Later, we will explain how to convert the general
problem to the homogeneous case.
As we described above, in this case, we set V = H01 (Ω).

Definition 5.8. We say that u ∈ H01 (Ω) is a weak solution to the homoge-
neous Dirichlet problem if
Z
B(u, v) = f (x)v(x) dx ∀ v ∈ H01 (Ω) .

5.2.2 The Neumann problem

Notice that the null space of L with u = 0 boundary conditions is trivial;


on the contrary, the null space of L with homogeneous Neumann boundary
conditions is one-dimensional and is spanned by the constant function u = 1.
One of the fundamental ideas in writing the variational form of the BVP
is to choose the functional framework in such a way as to mod-out any
nontrivial kernel. This “modding-out” coincides with the solvability criteria
of the Fredholm Alternative Theorem that we discuss in Section 5.7.
As such, for the homogeneous Neumann problem, we set V = H 1 (Ω)/R.

Definition 5.9. We say that u ∈ H 1 (Ω)/R is a weak solution to the Neu-


mann problem if
Z Z
B(u, v) = f (x)v(x) dx + g(x)v(x) dSx ∀ v ∈ H 1 (Ω)/R .
Ω ∂Ω

5.2.3 The Robin problem

For the homogeneous Robin problem, we observe that the kernel of L is once
again trivial. In this case we set V = H 1 (Ω).

Definition 5.10. We say that u ∈ H 1 (Ω) is a weak solution to the homo-


geneous Robin problem if
Z Z Z
B(u, v) + k(x)u(x)v(x) dSx = f (x)v(x) dx + g(x)v(x) dx
∂Ω Ω ∂Ω
∀ v ∈ H 1 (Ω) .
190 CHAPTER 5. Second Order Linear Elliptic Equations

5.2.4 A non-standard elliptic problem

We next consider the variational form of the elliptic BVP Lu = f in Ω with


the non-standard boundary condition aij u,j Ni − ∆g u = 0 on ∂Ω.
Following Definition 5.10, we see that integration by parts leads to the
following variational form
Z
B(u, v) − (∆g u)v dSx = (f, v)L2 (Ω) (5.13)
∂Ω

for all v in some space of test functions. Just as we used integration by parts
on the interior
Z elliptic operator, we can now do the same on the boundary
integral (∆g u)v dSx . In order to do so, we make use of a local coordinate
∂Ω
system (x1 , ..., xn−1 ) for ∂Ω, and note the surface measure is given by

p
dS = det g dx1 · · · dxn−1 .

It follows that
Z p Z
− (∆g u)v dSx = − [ det gg αβ u,β ],α v dx1 · · · dxn−1
∂Ω Z ∂Ω p
= g αβ u,β v,α det g dx1 · · · dxn−1
Z∂Ω

= g gradu(x), gradv(x) dSx ,
∂Ω

where gradu denotes the surface gradient of the function u. It follows that
we require both H 1 (Ω) regularity on the interior as well as H 1 (∂Ω) on the
boundary. Note that the kernel of L is once again spanned by the constant
functions.

This leads us to set V = H 1 (Ω)/R ∩ H 1 (∂Ω) = w ∈ H 1 (Ω)/R w ∈


H 1 (∂Ω) . We make the following


Definition 5.11. We say that u ∈ H 1 (Ω)/R ∩ H 1 (∂Ω) is a weak solution


to the non standard elliptic BVP if
Z Z

B(u, v) + g gradu(x), gradv(x) dSx = f (x)v(x) dx
∂Ω Ω
∀ v ∈ H 1 (Ω)/R ∩ H 1 (∂Ω) .
5.3. THE LAX-MILGRAM THEOREM 191

5.2.5 Linear elasticity


Dirichlet boundary conditions

The case of homogeneous Dirichlet boundary conditions for the vector prob-
lem is the same as for the scalar problem. In particular, we set V =
H01 (Ω; R3 ).

Definition 5.12. We say that u ∈ H01 (Ω; R3 ) is a weak solution to the


system of equations of linear elasticity with homogeneous Dirichlet boundary
conditions if

∂u` ∂v i
Z Z
C ijk` dx = f (x) · v(x) dx ∀ v ∈ H01 (Ω; R3 ) .
Ω ∂xk ∂xj Ω

Neumann boundary conditions

When we prescribe the traction vector on the boundary rather than the
displacement vector, the equation Div σ = 0 has a six-dimensional null space
spanned by the translations and rotations of R3 . In particular, the kernel is
spanned by α1 + α2 x where α1 is any constant vector and α2 is any skew-
symmetric matrix. Let S denote this span. Then we set V = H 1 (Ω; R3 )/S.

Definition 5.13. We say that u ∈ H 1 (Ω; R3 )/S is a weak solution to the


system of equations of linear elasticity with homogeneous Neumann boundary
conditions if
` i
ijk` ∂u ∂v
Z Z
C dx = f (x) · v(x) dx ∀ v ∈ H 1 (Ω; R3 )/S .
Ω ∂xk ∂xj Ω

5.3 The Lax-Milgram theorem

A general framework for proving the existence of a unique weak solution is


based on a generalization of the Riesz Representation theorem.
The objective is to find a unique weak solution given by Definition 5.7.
We restate our definition as follows: we suppose that V is a reflexive Banach
space, and that we have a continuous bilinear map B : V × V → R for which
we search for a unique element u ∈ V such that B(u, v) = hf, vi for all v ∈ V.
We are using the notation h·, ·i to denote the duality pairing between V and
its dual space V 0 , and we assume that the interior forcing function induces
192 CHAPTER 5. Second Order Linear Elliptic Equations

a continuous linear functional on V. For second-order elliptic BVPs, the


Banach space V will be a subspace of the Hilbert space H 1 (Ω) (or H 1 (Ω; Rn )
in the case of vector-valued solutions), and forcing function function will
typically have L2 (Ω) regularity.
The bilinear form B induces a linear operator L : V → V 0 defined by

Lu = B(u, ·) ∀u ∈ V or hLu, vi = B(u, v) ∀ u, v ∈ V .

Therefore, the variational formulation can be stated as

Lu = F ∈ V 0 ,

which means that


hLu, viV = hF, viV ∀v ∈ V.

Let B(V, V 0 ) denote the space of bounded linear operators from V to V0 .

Definition 5.14. L ∈ B(V, V 0 ) is coercive if there exists λ 0 > 0 such that

λ 0 kuk2V ≤ hLu, uiV ∀u ∈ V,

Theorem 5.15 (Lax-Milgram). If L ∈ B(V, V 0 ) and is coercive, then there


exists a unique solution to Lu = f ∈ V 0 and u ∈ V satisfies

1
kukV ≤ kf kV 0 . (5.14)
λ0

Proof. Step 1: (Uniqueness) Assume that v is also a solution to Lv = f .


Then
ku − vk2V ≤ λ−1
0 hL(u − v), u − viV = 0 .

Hence u = v in V.
Step 2: (Estimate) Assume that u ∈ V satisfies Lu = f ∈ V 0 . Then

kuk2V ≤ λ−1 −1 −1
0 hLu, uiV = λ 0 hf, uiV ≤ λ 0 kf kV 0 kukV ;

thus kukV ≤ λ−1


0 kf kV 0 .

Step 3: (Existence)
5.4. THE BBL INF-SUP CONDITION 193

1. We first prove that the range of L, R(L), is closed. Since R(L) is a


metric space, it suffices to show R(L) is complete. Let yn ∈ R(L) be
a Cauchy sequence; that is, given  > 0, there exists N > 0 so that
kyn − ym kV 0 ≤  whenever n, m ≥ N . Since each yn ∈ R(L), there
exists xn ∈ V such that Lxn = yn . Since L(xn − xm ) = yn − ym , by
(5.14),
kxn − xm kV ≤ λ−1
0  ∀ n, m ≥ N .

This implies xn is a Cauchy sequence in V which is a Banach space so


∃ x ∈ V such that xn → x ∈ V. Now

lim yn = lim Lxn = L lim xn = Lx ≡ y ∈ R(L) .


n→∞ n→∞ n→∞

So R(L) is closed.

2. Next, we show that R(L) = V 0 . Assume to the contrary that R(L) ⊆


V 0 . By the Hahn-Banach theorem, ∃ u ∈ (V 0 )0 = V (by reflexivity)
such that u|R(L) = 0 but hu, f i = 1. Then

kuk2V ≤ λ−1
0 hLu, uiV = 0

which contradicts the conclusion that u = 0. 

5.4 The BBL inf-sup condition3

The Lax-Milgram theorem is not sufficient to cover certain elliptic systems


of mixed type, such as a formulation of the Stokes equations with explicit
pressure and velocity. We shall describe a theory which generalizes the
Lax-Milgram theorem without assuming that L maps V into its dual space
V 0 . The replacement for the coercivity is the so-called inf-sup condition
introduced by Babuška, Brezzi, and Ladyzhenskaya.
Let V and W be Hilbert spaces with inner products (·, ·)V and (·, ·)W .
Let B : V × W → R be bilinear and continuous:

B(v, w) ≤ kBkkvkV kwkW ∀v ∈ V,w ∈ W.


3
Readers should skip this section on first reading.
194 CHAPTER 5. Second Order Linear Elliptic Equations

In particular, we could have V = W. Consider L : V → W 0 , L∗ : W → V 0


defined by

hLv, wiW = hL∗ w, viV = B(v, w) ∀v ∈ V,w ∈ W.

So L∗ is the adjoint operator associated with L. Note that by the continuity


of B, L and L∗ are both continuous.

Theorem 5.16 (Inf-Sup). The linear map L : V → W 0 is an isomorphism


if and only if the following conditions hold:
B(v, w)
(1) there exists α > 0 so that sup ≥ αkvkV ,
w∈W kwkW

(2) for every non-zero w ∈ W, there exists v ∈ V so that B(v, w) 6= 0.

Remark 5.17. The condition (1) is equivalent to

B(v, w)
inf sup ≥ α > 0. (5.15)
v∈V w∈W kvkV kwkW

and the condition (2) is equivalent to

sup B(v, w) > 0 ∀ w 6= 0 . (5.16)


v∈V

Note that (5.16) is also equivalent to ker L∗ = {0}.

Lemma 5.18 (The closed range theorem). Let X and Y be Banach spaces,
and L : X → Y be a continuous linear map. Then the following statements
are equivalent:

1. The image L(X) of X is closed in Y,

2. L(X) = (ker L∗ )0 , where for a closed subspace A ⊆ V∗ , A0 is the polar


set of A defined by

A0 ≡ w∗ ∈ V 0 hw∗ , ai = 0 ∀ a ∈ A .


Lemma 5.19 (Brezzi). The following statements are equivalent.


B(v, w)
(a) there exists α > 0 so that sup ≥ αkvkV ,
w∈W kwkW
§5.4 The BBL inf-sup condition 195

(b) L : V → W00 is an isomorphism and

kLvkW 0 ≥ αkvkV ∀v ∈ V,

(c) L∗ : W0⊥ → V 0 is an isomorphism and

kL∗ wkV 0 ≥ αkwkW ∀ w ∈ W⊥


0 ,

where W0 ≡ w ∈ W B(v, w) = 0 ∀ v ∈ V = ker L∗ .




Proof. Step 1: (a) ⇒ (b)


hLv, wi B(v, w)
kLvkW 0 = sup = sup ≥ αkvkV .
w∈W kwk W w∈W kwkW

This implies that L is one-to-one and so invertible in R(L) = L(V) which is


closed (same argument as the proof of the Lax-Milgram Theorem). More-
over, L−1 : R(L) → V is continuous. By the closed range theorem, L : V →
W00 is an isomorphism.
Step 2: (b) ⇒ (c)
Let w ∈ W⊥ ∗ 0
0 be arbitrary and consider map w ∈ W , z 7→ (z, w)W =
hw∗ , zi = w∗ (z). Since (z, w)W = 0 for all z ∈ W0 , we see that w∗ ∈ W00 .
By (i), there is a unique v ∈ W so that Lv = w∗ . Then

αkvkV ≤ kLvkW 0 = kw∗ kW 0 = kwkW .

Therefore,
hw∗ , wi B(v, w) hL∗ w, vi hL∗ w, vi
αkwkW = α =α =α ≤ .
kwkW kwkW kwkW kvkV
This implies L∗ is one-to-one, (L∗ )−1 : L∗ (W⊥ ⊥
0 ) → W0 is continuous,
and R(L∗ ) is closed. We then conclude by the closed graph theorem that
L∗ (W⊥ 0 0 0
0 ) = (ker L) = {0} = V .
Step 3: (c) ⇒ (a)
Take an arbitrary v ∈ V, and let w ∈ W⊥ ∗ ∗ 0
0 satisfy L w = v ∈ V with
kL∗ wkV 0 ≥ αkwkW . This implies
hv ∗ , vi hL∗ w, vi B(v, w)
kvkV = ∗
= ∗
≤ .
kv kV 0 kL wkV 0 αkwkW
Therefore,
B(v, w) B(v, w) B(v, w)
sup ≥ sup = sup ≥ αkvkV . 
w∈W kwk W w∈W⊥ kwk W v ∈V kwkW
∗ 0
0
196 CHAPTER 5. Second Order Linear Elliptic Equations

Proof of Theorem 5.16.

(⇐) Assume (1) and (2). Since (1) is the same as (a) of Lemma 5.19, (1)
⇔ (b) so that L : V → W00 is an isomorphism. It remains to prove
that W00 = W 0 . By (2), W0 = ker L∗ = {0} and hence W00 = W 0 .

(⇒) Suppose L : V → W 0 is an isomorphism. Given v ∈ V, let w∗ = Lv.


Since L−1 : W 0 → V is bounded, we find that
hLv, wi 1 1
sup = kLvkW 0 = kw∗ kW 0 ≥ kL−1 w∗ kV = kvkV .
w∈W kwkW C C
Therefore, α = 1/C will do the job for (1).

Moreover, L(V) = W 0 = (ker L∗ )0 , whence ker L∗ = {0} which implies


(2). 

Theorem 5.20 (Inf-Sup, another form). The following inf-sup condition is


equivalent to (1) and (2) in Theorem 5.16:
B(v, w) B(v, w)
inf sup = inf sup > 0. (5.17)
v∈V w∈W kvkV kwkW w∈V v∈W kvkV kwkW

B(v, w) B(v, w)
Proof. Let α = inf sup and β = inf sup . We first
kvkV kwkW
v∈V w∈W w∈W v∈V kvkV kwkW
show that (1) and (2) are equivalent to α > 0 and β > 0, and then prove
α = β.

(⇐) Note that (1) is the same as α > 0. It suffices to prove (1) and
(2) implies β > 0. Since (2) implies W0 = ker L∗ = {0}, W⊥
0 =
W. (5) (which is equivalent to (1)) then implies L∗ : W → V 0 is an
isomorphism with kL∗ wkV 0 ≥ αkwkW . Therefore,
B(v, w) 1 hL∗ w, vi 1
sup = sup = kL∗ wkV 0 ≥ α .
v∈V kvkV kwkW kwkW v∈V kvkV kwkW

(⇒) Suppose that α and β are both positive. β > 0 implies that ker L∗ =
{0} which is equivalent to (2). (1) is again the same as α > 0.

By Brezzi’s lemma, kL∗ wkV 0 ≥ αkwkW . Therefore,


B(v, w)
sup ≥α
v∈V kvkV kwkW
which implies β ≥ α. Since the argument is symmetric, α ≥ β as well and
hence α = β. 
5.5. EXISTENCE AND UNIQUENESS OF WEAK SOLUTIONS 197

Remark 5.21. We can write the condition α = β in a different form


kL−1 k = k(L∗ )−1 k for
1 1 kLvkW 0 B(v, w)
= = inf = inf sup
kL−1 k kvkV v∈V kvkV v∈V w∈W kvkV kwkW
sup
v∈V kLvk W 0

B(v, w) 1
= inf sup = .
w∈W v∈V kvkV kwkW k(L )−1 k

5.5 Existence and uniqueness of weak solutions

Basic Assumptions. In order to highlight the fundamental concepts, we


shall restrict our attention in this section to the case in which all of the lower-
order terms appearing in the operator L vanish: bi (x) = 0 for i = 1, ..., n
and c(x) = 0. In particular, we assume the following:
 
∂ ij ∂
1. L = − a (x) ;
∂xi ∂xj
2. L is uniformly elliptic in Ω;

3. the interior forcing function f is in L2 (Ω);

4. the boundary forcing g = 0 on ∂Ω.

The bilinear form B : V × V → R reduces to


∂u ∂v
Z
B(u, v) = aij (x) dx ,
Ω ∂xj ∂xi

and L : V ∩ v → (V ∩ v)0 is given by

hLu, vi = B(u, v) + b(u, v) ∀ u, v ∈ V ∩ v . (5.18)

5.5.1 A few Preliminary Lemmas

Lemma 5.22 (Poincaré inequality). Let Ω ⊆ Rn denote an open, bounded,


connected, and smooth domain. Then

ku − ūkL2 (Ω) ≤ CkDukL2 (Ω) ∀ u ∈ H 1 (Ω) , (5.19)


Z
where ū := − u(y)dy denotes the average value of u over Ω and the constant

C depends on Ω.
198 CHAPTER 5. Second Order Linear Elliptic Equations

Proof. Suppose for the sake of contradiction that (5.22) does not hold. Then
there is a sequence {uj }∞ 1
j=1 ⊆ H (Ω) satisfying

kuj − ūj kL2 (Ω) > jkDuj kL2 (Ω) , (5.20)

with an associated sequence on the unit ball of H 1 (Ω) given by


uj − ūj
wj = with kwj kL2 (Ω) = 1 and w̄j = 0 .
kuj − ūj kL2 (Ω)

According to (5.20), kDwj kL2 (Ω) < j −1 , so that kwj k2H 1 (Ω) < 1 + j −2 <
∞. Strong compactness, given by Theorem 2.109 (see also Theorem 2.176)
provides a subsequence {wj k } and a limit w ∈ L2 (Ω) such that wj k → w in
L2 (Ω) as k → ∞. The limit w satisfies w̄ = 0 and kwkL2 (Ω) = 1.
Letting φ ∈ C0∞ (Ω). We see that
Z Z
w(x)Dφ(x) dx = lim wj k (x)Dφ(x) dx
Ω k→∞ Ω
Z
= − lim Dwj k (x) φ(x) dx ≤ lim jk−1 kφkL2 (Ω) = 0 .
k→∞ Ω k→∞

This shows that the weak derivative of w exists and is equal to zero almost
everywhere, i.e. w ∈ H 1 (Ω) and Dw = 0 a.e. As Ω is connected, we see
that w is a constant, and since w̄ = 0, we see that w = 0, contradicting the
fact that kwkL2 (Ω) = 1. 

The identical proof also shows that the following two lemmas hold:

Lemma 5.23 (Poincaré inequality for H01 (Ω)). Let Ω ⊆ Rn denote an open,
bounded, connected, and smooth domain. Then

kukL2 (Ω) ≤ CkDukL2 (Ω) ∀ u ∈ H01 (Ω) , (5.21)

where the constant C depends on Ω.

Lemma 5.24 (Poincaré inequality for the Robin problem). Let Ω ⊆ Rn


denote an open, bounded, connected, and smooth domain. Then for k ∈
L∞ (∂Ω) and k ≥ 0 on ∂Ω and k > 0 on a set of surface measure greater
than zero. Then
 √ 
kukL2 (Ω) ≤ C k k uk2L2 (∂Ω) + kDukL2 (Ω) ∀ u ∈ H 1 (Ω) , (5.22)

where the constant C depends on Ω.


§5.5 Existence and uniqueness of weak solutions 199

Corollary 5.25. Whenever ū = 0, kDukL2 (Ω) is an equivalent norm on


H 1 (Ω). In particular, there exists constants C1 , C2 such that

C1 kDukL2 (Ω) ≤ kukH 1 (Ω) ≤ C2 kDukL2 (Ω) .

According to Corollory 5.25, kDukL2 (Ω) is an H 1 (Ω)-equivalent norm


on the spaces H01 (Ω) and H 1 (Ω)/R. Of course, the Poincaré inequality
also holds for vector-valued functions as well. Another important inequality
follows:

Lemma 5.26 (Korn’s first inequality). Suppose that Ω satisfies the assump-
tions of the Poincaré inequality. Then

kukH 1 (Ω) ≤ CkkL2 (Ω) ∀ u ∈ H01 (Ω; R3 ) ,

1 ∂ui ∂uj
where ij = ( + ) denotes the symmetric part of Du.
2 ∂xj ∂xi
Lemma 5.27 (Korn’s second inequality). Suppose that Ω satisfies the as-
sumptions of the Poincaré inequality. Then

kukH 1 (Ω) ≤ CkkL2 (Ω) ∀ u ∈ H 1 (Ω; Rn )/S .

The proofs of these two lemmas are similar to the proof the Poincaré
inequality, and are left to the reader.

Remark 5.28. The Korn inequalities show that control of the symmetric
part of the gradient of u in L2 (Ω) is sufficient to control the entire gradient
of u in L2 (Ω).

5.5.2 The Dirichlet problem

Theorem 5.29. For any f ∈ L2 (Ω), there exists a unique weak solution
u ∈ H01 (Ω) to the homogeneous Dirichlet problem
∂  ij ∂u 
− a =f in Ω , (5.23a)
∂xi ∂xj
u=0 on ∂Ω . (5.23b)

Moreover, the weak solution u satisfies

kukH 1 (Ω) ≤ Ckf kL2 (Ω) . (5.24)


200 CHAPTER 5. Second Order Linear Elliptic Equations

Proof. According to Definition 5.8, we must find a unique element u ∈


H01 (Ω) satisfying
Z
B(u, v) = f (x)v(x) dx ∀ v ∈ H01 (Ω) ,

or simply Lu = f in H −1 (Ω) with L defined (5.18) (in which b ≡ 0). The


proof follows by an application of the Lax-Milgram Theorem; hence, we
must verify the coercivity and boundedness assumptions of that theorem.
Step 1. Boundedness. By Corollary 5.25, kDukL2 (Ω) is an equivalent
norm on H01 (Ω). Thus, we must find a constant C > 0 such that

kLukH −1 (Ω) ≤ CkDukL2 (Ω) ∀ v ∈ H01 (Ω) .

By the definition of the dual space norm,

∂u ∂v
Z
kLukH −1 (Ω) = sup hLu, viH01 (Ω) = sup aij (x) dx
kvkH 1 (Ω) =1 kvkH 1 (Ω) =1 Ω ∂xj ∂xi
0 0

≤ kakL∞ (Ω) kDukL2 (Ω) ,

where kakL∞ (Ω) ≡ max kaij kL∞ (Ω) .


i,j
∂u
Step 2. Coercivity. Let ξi = or equivalently, ξ = Du. By uniform
∂xi
ellipticity,

∂u ∂u
aij = aij ξj ξi ≥ λ 0 |ξ|2 = λ 0 |Du|2 .
∂xj ∂xi

Since, by definition, hLu, ui = B(u, u), we see that

∂u ∂u
Z
B(u, u) = aij dx ≥ λ 0 kDuk2L2 (Ω) = λ 0 kuk2H 1 (Ω) ∀ u ∈ H01 (Ω) .
Ω ∂xj ∂xi 0

In other words, L : H01 (Ω) → H −1 (Ω) is coercive if L is uniformly elliptic.




Note the essential use of the Poincaré inequality to show that B(u, u)
is coercive with respect to H01 (Ω). The fact that u = 0 on ∂Ω, which in
turn gives a trivial kernel for the operator L, is the key to the coercivity
inequality.
§5.5 Existence and uniqueness of weak solutions 201

5.5.3 The Neumann problem

Theorem 5.30. For any f ∈ L2 (Ω), there exists a unique weak solution
u ∈ H 1 (Ω)/R to the homogeneous Neumann problem
∂  ij ∂u 
− a =f in Ω , (5.25a)
∂xi ∂xj
∂u
aij Ni = 0 on ∂Ω . (5.25b)
∂xj

Moreover, the weak solution u satisfies

kukH 1 (Ω) ≤ Ckf kL2 (Ω) . (5.26)

Proof. According to Definition 5.9, we must find a unique element u ∈


H 1 (Ω)/R satisfying
Z
B(u, v) = f (x)v(x) dx ∀ v ∈ H 1 (Ω)/R , (5.27)

or simply Lu = f in the dual space of H 1 (Ω)/R with L defined (5.18) (in


which b ≡ 0). Again we verify the coercivity and boundedness assumptions
for L.
Step 1. (Boundedness) We must find a constant C > 0 such that

kLukH −1 (Ω) ≤ CkukH 1 (Ω) ∀ u ∈ H 1 (Ω)/R ,

while the proof of the inequality above is the same as the case of Dirichlet
problem.
Step 2. (Coercivity) By Theorem 5.22, kDukL2 (Ω) is an equivalent norm
on H 1 (Ω)/R. Thus, by uniform ellipticity,
∂u ∂u
Z
B(u, u) = aij dx ≥ λ 0 kDuk2L2 (Ω) ∀ u ∈ H 1 (Ω)/R .
Ω ∂xj ∂xi

In other words, L : H 1 (Ω)/R → (H 1 (Ω)/R)0 is coercive if L is uniformly


elliptic. 

Remark 5.31. The use of the subspace H 1 (Ω)/R allows us to use the
Poincaré inequality and hence use kDukL2 (Ω) and an H 1 (Ω)-equivalent norm
on elements of H 1 (Ω)/R, and remove the one-dimensional kernel of the op-
erator L. Moreover, we can permit any forcing function f ∈ L2 (Ω). To see
202 CHAPTER 5. Second Order Linear Elliptic Equations

that this is indeed the case, note that both f and f − c, for a constant c,
induce the same linear functional on H 1 (Ω)/R. In Section 5.6, we will show
that the weak formulation (5.27) is equivalent to

∂u ∂v
Z Z
ij
a dx = (f − c)v dx ∀ v ∈ H 1 (Ω) , (5.28)
Ω ∂xj ∂xi Ω

Z
where c = f¯ = − f (x) dx. Using (5.28) as the variational formulation of

the homogeneous Neumann problem, the forcing function F = f −c is indeed
required to satisfy the solvability condition to guarantee the invertibility of
L.

5.5.4 The Robin problem

Theorem 5.32. If k ∈ L∞ (∂Ω) is non-negative on ∂Ω and is positive on a


portion of ∂Ω with positive measure, then for any f ∈ L2 (Ω), there exists a
unique weak solution u ∈ H 1 (Ω) to the Robin problem

∂  ij ∂u 
− a =f in Ω, (5.29a)
∂xi ∂xj
∂u
aij Ni + ku = 0 on ∂Ω . (5.29b)
∂xj

Moreover, the weak solution u satisfies

kukH 1 (Ω) ≤ Ckf kL2 (Ω) . (5.30)

Proof. According to Definition 5.10, we must find a unique element u ∈


H 1 (Ω) satisfying
Z
B(u, v) + b(u, v) = f (x)v(x) dx ∀ v ∈ H 1 (Ω) ,

Z
where b is defined by b(u, ϕ) = kuϕ dS. This is the same as finding a
∂Ω
solution to Lu = f in H 1 (Ω)0 with L defined by (5.18).
Step 1. (Boundedness of L) We must find a constant C > 0 such that

kLukH −1 (Ω) ≤ CkukH 1 (Ω) ∀ u ∈ H 1 (Ω) .


§5.5 Existence and uniqueness of weak solutions 203

By the Trace Theorem, a function w ∈ H 1 (Ω) belongs to L2 (∂Ω) and satis-


fies kwkL2 (∂Ω) ≤ kwkH 1 (Ω) . As a consequence,

kLukH 1 (Ω)0 = sup hLu, ϕiH 1 (Ω)


kϕkH 1 (Ω) =1

≤ kakL∞ (Ω) kDukL2 (Ω) + sup kkkL∞ (∂Ω) kukL2 (∂Ω) kϕkL2 (∂Ω)
kϕkH 1 (Ω) =1
 
≤ kakL∞ (Ω) + CkkkL∞ (∂Ω) kukH 1 (Ω) .

Step 2. (Coercivity of L) By uniform ellipticity,



B(u, u) + b(u, u) ≥ λ 0 kDuk2L2 (Ω) + k kuk2L2 (∂Ω) ∀ u ∈ H 1 (Ω) ,

so the coercivity of L amounts to the equivalency of the H 1 -norm and the


norm defined by

kwk ≡ k kwk2L2 (∂Ω) + kDwk2L2 (Ω) .

The equivalence of norms is guaranteed by the Poincaré inequality Lemma


5.24. Therefore, L : H 1 (Ω) → H 1 (Ω)0 is coercive if L is uniformly elliptic.


5.5.5 The non-standard elliptic problem

Theorem 5.33. Let ∂Ω be of class C 1 . Then for any f ∈ L2 (Ω), there


exists a unique weak solution u ∈ H 1 (Ω)/R ∩ H 1 (∂Ω) to the non-standard
elliptic problem
∂  ij ∂u 
Lu = − a =f in Ω, (5.31a)
∂xi ∂xj
∂u
aij Ni = ∆g u on ∂Ω . (5.31b)
∂xj

Moreover, the weak solution u satisfies

kukH 1 (Ω) + kukH 1 (∂Ω) ≤ Ckf kL2 (Ω) . (5.32)



Proof. Let V = H 1 (Ω)/R ∩ H 1 (∂Ω) ≡ w ∈ H 1 (Ω)/R w ∈ H 1 (∂Ω) .


According to Definition 5.10, we must find a unique element u ∈ V satisfying


Z
B(u, v) + b(u, v) = f (x)v(x) dx ∀v ∈ V ,

204 CHAPTER 5. Second Order Linear Elliptic Equations
Z
where b is defined by b(u, ϕ) = g(gradu, gradv)dS. This is the same as
∂Ω
finding a solution to Lu = f in V 0 with L defined (5.18).
Step 1. (Boundedness of L) We must find a constant C > 0 such that
h i
kLukV 0 ≤ C kukH 1 (Ω) + kukH 1 (∂Ω) ∀u ∈ V .

Since ∂Ω is of class C 1 , the metric tensor g ∈ L∞ (∂Ω). With kgkL∞ (∂Ω)


denoting max kgαβ kL∞ (∂Ω) , the Trace Theorem implies that
α,β

hLu, viV ≤ kakL∞ (Ω) kukH 1 (Ω) kϕkH 1 (Ω) + kgkL∞ (∂Ω) kukH 1 (∂Ω) kϕkH 1 (∂Ω)
h i
≤ kakL∞ (Ω) + kgkL∞ (∂Ω) kukV kϕkV ;

thus kLukV 0 ≤ CkukV for all u ∈ V.


Step 2. (Coercivity of L) By uniform ellipticity and the Poincaré in-
equality, for all u ∈ V,
1
hLu, uiV ≥ λ 0 kDuk2L2 (Ω) + kgraduk2L2 (∂Ω) ≥ kuk2H 1 (Ω) + kgraduk2L2 (∂Ω)
C1
for some constant C1 > 0. Moreover, the Trace Theorem implies that for
some constant C2 > 0, kukL2 (∂Ω) ≤ C2 kukH 1 (Ω) ; thus the inequality above
further implies that for some constant C > 0,
1h i
hLu, uiV ≥ kuk2H 1 (Ω) + kuk2H 1 (∂Ω) ∀u ∈ V
C
which guarantees the coercivity of L. 

5.5.6 The elasticity equations


Dirichlet boundary conditions

Theorem 5.34. If the elasticity tensor C ijk` ∈ L∞ (Ω) satisfies the positivity
and symmetry assumptions of Definition 5.3, then for any f ∈ L2 (Ω; R3 ),
there exists a unique weak solution u ∈ H01 (Ω; R3 ) to the linear elasticity
equation
∂  ijk` ∂u` 
− C = fi in Ω, (5.33a)
∂xj ∂xk
u=0 on ∂Ω . (5.33b)

Moreover, the weak solution u satisfies

kukH 1 (Ω;R3 ) ≤ Ckf kL2 (Ω;R3 ) .


§5.5 Existence and uniqueness of weak solutions 205

Proof. According to Definition 5.12, we must find a unique element u ∈


H01 (Ω; R3 ) satisfying
∂u` ∂v i
Z Z
C ijk` dx = f (x) · v(x) dx ∀ v ∈ H01 (Ω; R3 ) .
Ω ∂xk ∂xj Ω

This is the same as finding a solution to Lu = f in H −1 (Ω; R3 ) if L is defined


by
∂u` ∂v i
Z
hLu, viH01 (Ω;R3 ) = C ijk` dx ∀ u, v ∈ H01 (Ω; R3 ) .
Ω ∂xk ∂xj
As before, we verify the coercivity and boundedness assumptions of L.
Step 1. (Boundedness of L) We must find a constant C > 0 such that

kLukH −1 (Ω;R3 ) ≤ CkukH 1 (Ω;R3 ) ∀ u ∈ H 1 (Ω; R3 ) .

Nevertheless, it follows from C ijk` ∈ L∞ (Ω) that

kLukH −1 (Ω;R3 ) ≤ max kC ijk` kL∞ (Ω) kDukL2 (Ω;R3 ) ∀ u ∈ H01 (Ω; R3 ) .
i,j,k,`

Step 2. (Coercivity of L) We note that the symmetry of C ijk` implies


that

C ijk` ui ,j u` ,k = C ijk` ij k` ;

thus the positivity of C ijk` and Korn’s first inequality implies that
1
hLu, uiH01 (Ω;R3 ) ≥ λkk2L2 (Ω) ≥ kuk2H 1 (Ω;R3 ) ∀ u ∈ H01 (Ω; R3 ) ;
C 0

so we establish the coercivity of L. Therefore, L : H01 (Ω; R3 ) → H −1 (Ω; R3 )


is coercive if the elasticity tensor C ijk` satisfies the positivity and the sym-
metry assumptions of Definition 5.3. 

Neumann boundary conditions

Theorem 5.35. If the elasticity tensor C ijk` ∈ L∞ (Ω) satisfies the positivity
and symmetry assumptions of Definition 5.3, then for any f ∈ L2 (Ω; R3 ),
there exists a unique weak solution u ∈ H 1 (Ω; R3 )/S to the linear elasticity
equation
∂  ijk` ∂u` 
− C = fi in Ω, (5.34a)
∂xj ∂xk
∂u`
C ijk` Nj = 0 on ∂Ω . (5.34b)
∂xk
206 CHAPTER 5. Second Order Linear Elliptic Equations

Moreover, the weak solution u satisfies

kukH 1 (Ω;R3 ) ≤ Ckf kL2 (Ω;R3 ) .

Proof. The proof for the case of the Neumann boundary condition is almost
the same as the Dirichlet case except that the coercivity of L is guaranteed
by Korn’s second inequality which implies that
1
hLu, uiH 1 (Ω;R3 ) ≥ kuk2H 1 (Ω;R3 ) ∀ u ∈ H 1 (Ω; R3 )/S. 
C

5.6 Weak solutions with distributional forcing func-


tions

Recall that (·, ·)H denotes the inner product of the Hilbert space H, and
h·, ·iX denotes the duality pairing between X and its dual space X 0 .
When the interior forcing function f ∈ L2 (Ω) as in the previous section,
the operator L maps from a Hilbert space V into a proper subspace of the
dual space V 0 . Notice, however, that the Lax-Milgram Theorem allows us
to establish existence and uniqueness for elliptic BVPS when the forcing
function is given in the dual space V 0 of the solution space V.
In order to consider distributional forcing functions which are elements of
V 0 , it is necessary to explain the representation theory for these various dual
spaces. In the following example, we provide some representations of V 0 for
a variety of different boundary conditions that we have already discussed.

Example 5.36.

1. Let H1 = H01 (Ω), and f ∈ H −1 (Ω) = H01 (Ω)0 . By the Riesz represen-
tation theorem there exists a unique u ∈ H01 (Ω) such that
Z
hf, ϕiH1 = (Du, Dϕ)L2 (Ω) = Du · Dϕ dx ∀ ϕ ∈ H1 .

2. Let H2 = H 1 (Ω), and f ∈ H 1 (Ω)0 . By the Riesz representation theo-


rem there exists an u ∈ H 1 (Ω) such that
Z
hf, ϕiH2 = (u, ϕ)H 1 (Ω) = (uϕ + Du · Dϕ) dx ∀ ϕ ∈ H2 .

§5.6 Weak solutions with distributional forcing functions 207

n Z o
3. Let H3 = H (Ω)/R ≡ w ∈ H 1 (Ω)
1
w(x) dx = 0 be endowed with


the usual H 1 -norm, that is, kwkH3 = kwkH 1 (Ω) . Let f ∈ H30 , then by
the Riesz representation theorem, there exists an u ∈ H 1 (Ω)/R such
that
Z

hf, ψiH3 = (u, ψ)L2 (Ω) = uψ + Du · Dψ dx ∀ ψ ∈ H3 .

(5.35)
On the other hand, by the Poincaré inequality, (Du, Dψ)L2 (Ω) is an
equivalent inner product on H3 . Using this as the inner product on
H3 , for any f ∈ H30 , there exists v ∈ H 1 (Ω)/R such that

hf, ψiH3 = (Dv, Dψ)L2 (Ω) ∀ ψ ∈ H3 . (5.36)

4. Let f ∈ H 1 (Ω)0 , then f ∈ H30 ; thus there exists u ∈ H3 such that

hf, ψiH 1 (Ω) = (u, ψ)L2 (Ω) + (Du, Dψ)L2 (Ω) ∀ ψ ∈ H 1 (Ω)/R ,

where hf, ψiH 1 (Ω)/R = hf, ψiH 1 (Ω) is used to obtain the second equality.
In particular, for any ϕ ∈ H 1 (Ω), ψ ≡ ϕ − ϕ̄ ∈ H 1 (Ω)/R can be used
in the equality above, thus by (u, 1)L2 (Ω) = 0,

∀ ϕ ∈ H 1 (Ω) .


f, ϕ − ϕ̄ H 1 (Ω) = (u, ϕ)L2 (Ω) + (Du, Dϕ)L2 (Ω)

Let f ∈ C ∞ (Ω) and f → f in H 1 (Ω)0 . Then


Fubini’s Theorem
lim (f¯ , ϕ)L2 (Ω)


f, ϕ̄ H 1 (Ω)
= lim f , ϕ̄ L2 (Ω) =
→0 →0

= hf, 1iH 1 (Ω) , ϕ L2 (Ω) ;

thus u satisfies

f − hf, 1iH 1 (Ω) , ϕ H 1 (Ω) = (u, ϕ)L2 (Ω) + (Du, Dϕ)L2 (Ω) ∀ ϕ ∈ H 1 (Ω) ,

or if c = hf, 1iH 1 (Ω) ,

∀ ϕ ∈ H 1 (Ω) .

hf, ϕiH 1 (Ω) = (u + c, ϕ)L2 (Ω) + D(u + c), Dϕ L2 (Ω)

In other words, u + hf, 1iH 1 (Ω) is the representation of f ∈ H 1 (Ω)0 .


Moreover, even though in general f and f − hf, 1iH 1 (Ω) are different
bounded linear functionals on H 1 (Ω), they are the same in the dual
space of H 1 (Ω)/R.
208 CHAPTER 5. Second Order Linear Elliptic Equations

5. Let H4 = H01 (Ω; Rn ), that is, u ∈ H4 if and only if u = (u1 , · · · , un )


and ui ∈ H01 (Ω) for i = 1, · · · , n. Then H4 is a Hilbert space endowed
n n
(ui , v i )H1 = (Dui , Dv i )L2 (Ω) .
P P
with the inner product (u, v)H4 =
i=1 i=1
If f = (f 1 , · · · , f n ) ∈ H −1 (Ω; Rn ) ≡ H01 (Ω; Rn )0 , by the Riesz repre-
sentation theorem, there exists a u = (u1 , · · · , un ) ∈ H01 (Ω; Rn ) such
that
n
X
hf, ϕiH4 = (u, ϕ)H4 = (Dui , Dϕ i )L2 (Ω) ∀ ϕ ∈ H4 .
i=1

1 (Ω; Rn ), that is, u ∈ H if and only if


6. Let H5 = Hdiv 5

(a) u ∈ Rn , u = (u1 , · · · , un ), ui ∈ H 1 (Ω) for i = 1, · · · , n,


n
ui ,i = 0 in Ω, and
P
(b) divu ≡
i=1
(c) u · N = 0 on ∂Ω.

H5 is a Hilbert space endowed with the inner product


n
X
(u, ϕ)H5 = (ui , ϕ i )H 1 (Ω)
i=1
n
X
(ui , ϕ i )L2 (Ω) + (Dui , Dϕ i )L2 (Ω)
 
= ∀ ϕ ∈ H5 .
i=1

1 (Ω; Rn )0 , by the Riesz representation theo-


If f = (f 1 , · · · , f n ) ∈ Hdiv
1 (Ω; Rn ) such that
rem, there exists a u = (u1 , · · · , un ) ∈ Hdiv
n
X
(ui , ϕ i )L2 (Ω) + (Dui , Dϕ i )L2 (Ω)
 
hf, ϕiH5 = ∀ ϕ ∈ H5 .
i=1

Example 5.36 gives some of the representation formulas for distributional


forcing functions. In the case that n = 1, a simple example is provided by

f =δ

where δ denotes the Dirac delta distribution. For example, with n = 1,


H01 (Ω) ⊆ C 0 (Ω) with continuous embedding, and hence C 0 (Ω)0 ⊆ H −1 (Ω),
so δ ∈ H −1 (Ω).
On the other hand, when f is a function in L2 (Ω), the above represen-
tations are not needed, as we discuss in the following remark.
§5.6 Weak solutions with distributional forcing functions 209

Remark 5.37. Let Hi be the Hilbert spaces in Example 5.36, and suppose
that a square integrable function f is an element in Hi0 . It follows that the
action of f on ϕ ∈ Hi is defined by hf, ϕiHi = (f, ϕ)L2 (Ω) . In particular a
function f ∈ L2 (Ω) defines a distribution Tf ∈ Hi0 by the formula

hTf , ϕiHi = (f, ϕ)L2 (Ω) .

With a slight abuse of notation, we will represent both functions f ∈ L2 (Ω)


and the induced distribution Tf by the symbol f .

With the notion of distributional forcing, we have the following theorems


which generalize Theorem 5.29, 5.30, 5.32, 5.33, 5.34 and 5.35.

Theorem 5.38. For any f ∈ H −1 (Ω), there exists a unique weak solution
u ∈ H01 (Ω) to the homogeneous Dirichlet problem (5.23) satisfying

B(u, v) = hf, viH01 (Ω) ∀ v ∈ H01 (Ω) ,

Moreover, the weak solution u satisfies

kukH 1 (Ω) ≤ Ckf kH −1 (Ω) . (5.37)

Theorem 5.39. For any f in the dual space of H 1 (Ω)/R, there exists a
unique solution u ∈ H 1 (Ω)/R to

B(u, v) = hf, viH 1 (Ω)/R ∀ v ∈ H 1 (Ω)/R ,

Moreover, the weak solution u satisfies

kukH 1 (Ω) ≤ Ckf k(H 1 (Ω)/R)0 . (5.38)

In particular, for any f ∈ H 1 (Ω)0 satisfying hf, 1iH 1 (Ω) = 0, there exists a
unique solution u ∈ H 1 (Ω)/R to

B(u, v) = hf, viH 1 (Ω)0 ∀ v ∈ H 1 (Ω)

satisfying kukH 1 (Ω) ≤ Ckf kH 1 (Ω)0 .


210 CHAPTER 5. Second Order Linear Elliptic Equations

Theorem 5.40. If k ∈ L∞ (∂Ω) is non-negative on ∂Ω and is positive on


a portion of ∂Ω, then for any f ∈ H 1 (Ω)0 , there exists a unique solution
u ∈ H 1 (Ω) to
Z
B(u, v) + kuvdS = hf, viH 1 (Ω) ∀ v ∈ H 1 (Ω) .
∂Ω

Moreover, the weak solution u satisfies

kukH 1 (Ω) ≤ Ckf kH 1 (Ω)0 . (5.39)

Theorem 5.41. Let ∂Ω be of class C 1 . Then for any f in the dual space of
H 1 (Ω)/R ∩ H 1 (∂Ω), there exists a unique solution u ∈ H 1 (Ω)/R ∩ H 1 (∂Ω)
to
Z
B(u, v) + g(gradu, gradv) dS = hf , viH 1 (Ω)/R∩H 1 (∂Ω)
∂Ω
∀ v ∈ H 1 (Ω)/R ∩ H 1 (∂Ω) .

Moreover, the weak solution u satisfies

kukH 1 (Ω) + kukH 1 (∂Ω) ≤ Ckf k(H 1 (Ω)/R∩H 1 (∂Ω))0 . (5.40)

Theorem 5.42. If the elasticity tensor C ijk` ∈ L∞ (Ω) satisfies the positivity
and symmetry assumptions of Definition 5.3, then for any f ∈ H −1 (Ω; R3 ),
there exists a unique weak solution u ∈ H01 (Ω; R3 ) to the linear elasticity
equation (5.33) satisfying
Z
C ijk` u` ,k v i ,j dx = hf, viH01 (Ω;R3 ) ∀ v ∈ H01 (Ω; R3 ) .

Moreover, the weak solution u satisfies

kukH 1 (Ω;R3 ) ≤ Ckf kH −1 (Ω;R3 ) .

Theorem 5.43. If the elasticity tensor C ijk` ∈ L∞ (Ω) satisfies the positivity
and symmetry assumptions of Definition 5.3, then for any f ∈ L2 (Ω; R3 ),
there exists a unique solution u ∈ H 1 (Ω; R3 )/S to
Z
C ijk` u` ,k v i ,j dx = hf, viH 1 (Ω;R3 )/S ∀ v ∈ H01 (Ω; R3 )/S .

Moreover, the weak solution u satisfies

kukH 1 (Ω;R3 ) ≤ Ckf k(H 1 (Ω;R3 )/S)0 .


5.7. GENERAL SECOND ORDER ELLIPTIC OPERATORS 211

5.7 General second order elliptic operators

Suppose now we want to solve


∂  ij ∂u  ∂u
Lu = − a + bi + cu = f in Ω, (5.41a)
∂xi ∂xj ∂xi
u=0 on ∂Ω , (5.41b)

where L is strictly elliptic, a, b, c ∈ L∞ (Ω). Let L : H01 (Ω) → H −1 (Ω) be


defined by for ϕ ∈ H01 (Ω),
ij ∂v ∂ϕ i ∂v
Z Z Z
hLv, ϕiH01 (Ω) = a dx + b ϕdx + cvϕ dx .
Ω ∂xj ∂xi Ω ∂xi Ω

It is easy to see that L : H01 (Ω) → H −1 (Ω) is continuous. In fact,

kLukH −1 (Ω) = sup hLu, ϕiH01 (Ω)


kϕkH 1 (Ω) =1
0
h i
≤ max kaij kL∞ (Ω) + kbkL∞ (Ω) + kckL∞ (Ω) kukH01 (Ω) .
i,j

However, it is not clear that if there exists γ0 > 0 so that γkuk2H 1 (Ω) ≤
0
hLu, uiH01 (Ω) . In fact, if c  0, L fails to be coercive. Let I : H01 (Ω) →
H −1 (Ω) be defined by
Z
hIv, ϕiH01 (Ω) = vϕ dx .

Then we have the following

Lemma 5.44. For some γ > 0 sufficient large, (L + γI) : H01 (Ω) → H −1 (Ω)
is coercive.

Proof. By the positivity of aij ,


∂u ∂u
Z
λ 0 kDuk2L2 (Ω) ≤ aij dx
Ω ∂xj ∂xi
Z Z
= hLu, uiH01 (Ω) − (b · Du)udx − cu2 dx
Ω Ω
≤ hLu, uiH01 (Ω) + kbkL∞ (Ω) kDukL2 (Ω) kukL2 (Ω) + kckL∞ (Ω) kuk2L2 (Ω)
λ0 h 1 i
≤ hLu, uiH01 (Ω) + kuk2H 1 (Ω) + kbk2L∞ (Ω) + kckL∞ (Ω) kuk2L2 (Ω) ,
2 0 2λ 0
where we use Young’s inequality
1 2
ab ≤ a2 + b ∀ > 0
4
212 CHAPTER 5. Second Order Linear Elliptic Equations

1
to obtain the last inequality. Choose γ > kbk2L∞ (Ω) + kckL∞ (Ω) to con-
2λ 0
clude that
λ0
kuk2H 1 (Ω) ≤ hLu, uiH01 (Ω) + γkuk2L2 (Ω) = hLu + γIu, uiH01 (Ω) . 
2 0

Theorem 5.45. For all f ∈ H −1 (Ω), there exists unique solution u ∈


H01 (Ω) to (L + γI)u = f in H −1 (Ω) if γ is large enough (independent of f ).

From now on, the solution u ∈ H01 (Ω) to (L + γI)u = f in H −1 (Ω)


is denoted by (L + γI)−1 f . The goal for the remaining section is to find
conditions of f ∈ L2 (Ω) under which Lu = f have a unique solution.

Definition 5.46. We call L∗ the formal adjoint of L and define it by

(L∗ v, u)L2 (Ω) = (Lu, v)L2 (Ω) ∀ u, v ∈ H01 (Ω) .

Integrating by parts, since u, v vanish on the boundary,

ij ∂u ∂v i ∂u
Z Z Z
(Lu, v)L2 (Ω) = a dx + b vdx + cuvdx
Ω ∂xj ∂xi Ω ∂xi Ω
∂  ij ∂v  ∂
Z Z Z
i
= − a udx + u (−b v)dx + cuvdx .
Ω ∂xj ∂xi Ω ∂xi Ω

Therefore,
∂  ij ∂v  ∂ i
L∗ v = − a − (b v) + cv .
∂xj ∂xi ∂xi

Theorem 5.47 (Fredholm Alternative). L2 (Ω) = N (L∗ ) ⊕ R(L), where


n o
N (L∗ ) = v ∈ H01 (Ω) L∗ v = 0 .

So either there exists a unique solution to Lu = f ∈ L2 (Ω) (the case


N (L∗ ) = {0}) or there is a v 6= 0 such that L∗ v = 0.

Theorem 5.47 follows from the theory of compact operator: for compact
operator K : H → H, H Hilbert (see Appendix B.3 for details),

(1) K ∗ is also compact;

(2) dim N (I − K) = dim(N (I − K ∗ )) < ∞;

(3) R(I − K) = R(Id − K ∗ ) is closed;


§5.7 General second order elliptic operators 213

(4) N (I − K ∗ )⊥ = R(I − K);

(5) (I − K)u = h has a unique solution in H or ∃ v 6= 0 in H solving


(I − K ∗ )v = 0.

Theorem 5.48 (Fredholm Alternative: another version). For all f ∈ L2 (Ω),


there exists a unique solution u ∈ H01 (Ω) to Lu = f or N (L∗ ) is not trivial.

Proof. Lu = f has a weak solution if and only if Lu + γu = f + γu has a


weak solution. By Theorem 5.45, we may choose γ > 0 so that (L + γI)−1
exists and

u = (L + γI)−1 f + γ(L + γI)−1 u or [I − γ(L + γI)−1 ]u = (L + γI)−1 f .

Claim: γ(L + γI)−1 : L2 (Ω) → L2 (Ω) is compact.

Again by Theorem 5.45, for all g ∈ L2 (Ω), there exists a unique solution u
to (L + γI)u = g and

kukH01 (Ω) = k(L + γI)−1 gkH01 (Ω) ≤ CkgkL2 (Ω) .

So (L + γI)−1 is a bounded linear operator from L2 (Ω) → H01 (Ω). By


Rellich’s theorem, H01 (Ω) is compactly embedded into L2 (Ω), so γ(L+γI)−1 :
L2 (Ω) → L2 (Ω) is compact. Therefore, (5) implies that it suffices to show
that
∃ v 6= 0 3 I − γ(L + γI)−1∗ v = 0 ⇔ {0} ( N (L∗ ) .


However, I − γ(L + γI)−1∗ v = 0 ⇔ (L + γI)∗ v = γv ⇔ L∗ v = 0.





Remark 5.49.

1. By fact (2), the last line of the proof also shows that dim(N (L∗ )) =
dim(N (L)) < ∞.

2. By fact (4), Lu = f has a solution (or equivalently, u solves (I −


γ(L + γI)−1 )u = (L + γI)−1 f ) if (L + γI)−1 f is orthogonal to v for all
v ∈ N (I − γ(L + γI)−1∗ ), i.e., ((L + γI)−1 f, v)L2 (Ω) = 0 for all such v.
But

0 = (L + γI)−1 f, v = f, (L + γI)−1∗ v
 
L2 (Ω) L2 (Ω)
= γ(f, v)L2 (Ω) .

What we have established is Fredholm Alternative theorem L2 (Ω) =


R(L) ⊕ N (L∗ ).
214 CHAPTER 5. Second Order Linear Elliptic Equations

3. Recall we have proved this for open bounded Ω. For unbounded Ω, the
Fredholm Alternative is not necessary true due to failure of Rellich’s
theorem.

4. In the proof of the Fredholm Alternative, the use of Dom(L) = H01 (Ω)
is not essential: it can be replaced by other spaces (but it might be dif-
ficult to compute the formal adjoint L∗ ). The same proof goes through
if Rellich’s theorem (or compactness arguments) can still be applied.

Theorem 5.50. Suppose that f ∈ L2 (Ω).

(1) If N (L∗ ) = {0}, then there exists a unique weak solution u ∈ H01 (Ω)
to Lu = f in H −1 (Ω). Furthermore,

kukH 1 (Ω) ≤ Ckf kL2 (Ω) . (5.42)

(2) If N (L∗ ) 6= {0} and f ⊥ N (L∗ ), then there exist multiple weak solu-
tions in H01 (Ω) to Lu = f , and each solution u satisfies
h i
kukH 1 (Ω) ≤ C kukL2 (Ω) + kf kL2 (Ω) . (5.43)

(3) If N (L∗ ) 6= {0} and f 6⊥ N (L∗ ), then there exists no weak solution in
H01 (Ω) to Lu = f .

Proof. (3) is proved in the previous Remark. Now suppose f ⊥ N (L∗ ) and
u is a weak solution to Lu = f . Then
Z Z Z Z
aij u,j v,i dx + bi u,i v dx + cuv dx = f v dx .
Ω Ω Ω Ω

Let v = u. Then

λ 0 kDuk2L2 (Ω)
1
≤ kbkL∞ (Ω) kDukL2 (Ω) kukL2 (Ω) + kckL∞ (Ω) kuk2L2 (Ω) + kf k2L2 (Ω) + kuk2L2 (Ω)

2
λ0 h 1 1 i 1
≤ kDuk2L2 (Ω) + kbk2L∞ (Ω) + kckL∞ (Ω) + kuk2L2 (Ω) + kf k2L2 (Ω) .
2 2λ 0 2 2
Therefore, by Poincaré’s inequality,
h i
kuk2H 1 (Ω) ≤ CkDuk2L2 (Ω) ≤ C kuk2L2 (Ω) + kf k2L2 (Ω) .
5.8. HORIZONTAL CONVOLUTION-BY-LAYERS AND COMMUTATION ESTIMATES215

This proves (2).


Now suppose that N (L∗ ) = {0}. Then fact (4) suggests that R(I −
γ(γL + γI)−1 ) = L2 (Ω); hence by the bounded inverse theorem B.14, the
inverse of I − γ(γL + γI)−1 belongs to B(L2 (Ω), L2 (Ω)). Since u = I −

−1
γ(L + γI)−1 (L + γI)−1 f ,

kukL2 (Ω) ≤ Ck(L + γI)−1 f kL2 (Ω) ≤ Ckf kL2 (Ω) .

Estimate (5.42) then follows from (5.43) and the L2 -estimate above. 

5.8 Horizontal convolution-by-layers and commu-


tation estimates
5.8.1 Horizontal convolution-by-layers
 
With xh = (x1 , · · · , xn−1 ), we define ρ ∈ C0∞ (Rn−1 ) by ρ(xh ) = C exp 1
|xh |2 −1
ifZ |xh | < 1 and ρ(xh ) = 0 if |xh | ≥ 1; we then select the constant C so that
ρ dxh = 1. We define ρ (xh ) = −(n−1) ρ(−1 xh ). It follows that for
Rn−1
 > 0, 0 ≤ ρ ∈ C0∞ (R2 ) with spt(ρ ) ⊆ B(0, ).
We define the horizontal convolution-by-layers operator Λ as follows:
Z
Λ f (xh , xn ) = ρ (xh − yh )f (yh , xn )dyh for f (·, xn ) ∈ L1 (Rn−1 ) .
Rn−1

It should be clear that Λ smooths functions defined on Rn along all hor-


izontal subspaces, but does not smooth these function in the vertical xn -
direction. On the other hand, we can restrict the operator Λ to act on
functions f : Rn−1 → R as well, in which case Λ becomes the usual mollifi-
cation operator.
By standard properties of convolution, there exists a constant C which
is independent of , such that for s ≥ 0,

kΛ F kH s (Ω) ≤ CkF kH s (Ω) ∀ F ∈ H s (Ω) ,

and
kΛ F kH s (∂Ω) ≤ CkF kH s (∂Ω) ∀ F ∈ H s (∂Ω) .

Furthermore,

¯  F kL2 (Ω) ≤ CkF kL2 (Ω)


k∂Λ ∀ F ∈ L2 (Ω) , (5.44)
216 CHAPTER 5. Second Order Linear Elliptic Equations

where ∂¯ denotes the vector of horizontal partial derivatives:


 
¯ ∂ ∂
∂= , · · ·, .
∂x1 ∂xn−1

5.8.2 Commutation estimates

Definition 5.51. Let X be a Banach space, and A : X → X and B : X → X


 
be two operators. The commutator of A and B, denoted by A, B , is defined
by
 
A, B w = A(Bw) − B(Aw) ∀w ∈ X . (5.45)

Lemma 5.52. For f ∈ W 1,∞ (Rn+ ) and g ∈ L2 (Rn+ ), there is a generic


constant C independent of  such that

∂¯ Λ , f g 2 n = ∂¯ Λ (f g) − f Λ g 2 n
    
L (R+ ) L (R+ )

≤ C f 1,∞ n g 2 n
W (R+ ) L (R+ )
(5.46)

for all  > 0.

Proof. By definition,
Z
   
Λ , f g(x) = ρ (xh − yh ) f (yh , xn ) − f (xh , xn ) g(yh , xn ) dyh .
B(xh ,)

By Morrey’s inequality, f (yh , xn ) − f (xh , xn ) ≤ Ckf kW 1,∞ (Rn+ ) . Since

¯  (xh − yh ) = 1 (Dρ) xh − yh ,
 
∂ρ
n 
we see that
Z
1  x − y 
∂¯ Λ , f g (x) ≤ Ckf kW 1,∞ (Rn )
   h h
(Dρ) g(yh , xn ) dyh .
+
Rn−1 n−1 
Finally, we note that
Z
1 x − y  Z
h h
n−1
|Dρ| dyh = |Dρ|(zh )dzh = kDρkL1 (Rn−1 ) ;
Rn−1   Rn−1

thus by Young’s inequality,

∂¯ Λ , f g (·, xn ) 2 n−1
  
L (R )
1  x − · 
h
≤ Ckf kW 1,∞ (Rn+ ) n−1 (Dρ) 1 n−1 kg(·, xn )kL2 (Rn−1 )

  L (R )
≤ Ckf kW 1,∞ (Rn+ ) kg(·, xn )kL2 (Rn−1 ) .
5.9. ELLIPTIC REGULARITY 217

The desired estimate (5.46) then follows from integrating the square of the
inequality above in xn over the interval (0, ∞). 

The same proof leads to the following

Lemma 5.53. For f ∈ W 1,∞ (Ω) and g ∈ L2 (Ω) with compact support,
there is a generic constant C independent of  such that
    
D η ∗, f g 2 = D η ∗ (f g) − f η ∗ g 2
L (Ω) L (Ω)

≤ Ckf kW 1,∞ (Ω) kgkL2 (Ω) (5.47)


  
for all  < min dist ∂Ω, spt(f ) , dist ∂Ω, spt(g) .

5.9 Elliptic regularity


5.9.1 Dirichlet problem with homogeneous boundary condi-
tion

Theorem 5.54. Suppose that Ω ⊆ Rn is an open, bounded domain with


smooth boundary, L is uniformly elliptic, and aij ∈ C 1 (Ω). Then for any
f ∈ L2 (Ω), the unique weak solution u to the Dirichlet problem (5.23) in
fact belongs to H 2 (Ω) ∩ H01 (Ω) and satisfies

kukH 2 (Ω) ≤ Ckf kL2 (Ω) , (5.48)

where C depends on kaij kC k (Ω) and the regularity of ∂Ω.

Proof. The goal is to show that the function u ∈ H01 (Ω) satisfying
Z Z
ij
a u,j v,i dx = f vdx ∀ v ∈ H01 (Ω) (5.49)
Ω Ω

satisfies (5.48). We divide the proof into several steps as follows:


Step 1. Interior estimates. We show that for any given Ω0 ⊂⊂Ω, the weak
solution u ∈ H 2 (Ω0 ) and satisfies

kukH 2 (Ω0 ) ≤ Ckf kL2 (Ω) , (5.50)

where C depends on Ω0 and in general C → ∞ as dist(∂Ω, Ω0 ) → 0. Interior


estimates are performed away from the boundary using cut-off functions
218 CHAPTER 5. Second Order Linear Elliptic Equations

with compact support. Consequently, we are permitted to use integration


by parts in all directions to obtain such estimates.
Step 2. Estimates near the boundary. Estimates in regions near the bound-
ary are significantly more delicate. In such regions, we remap the local neigh-
borhood of the boundary into the upper-half of the unit sphere, and perform
integration by parts only in horizontal directions – the normal derivatives
are then estimated using the structure of the elliptic operator, and in small
open neighborhood U of ∂Ω, we find that

kukH 2 (U ) ≤ Ckf kL2 (Ω) (5.51)

The combination of (5.50) and (5.51) then leads to (5.48).


Step 1. Interior estimates.
Choose any (nonempty) open sets Ω1 ⊂⊂Ω2 ⊂⊂Ω and let ζ ∈ C0∞ (Ω2 )
with 0 ≤ ζ ≤ 1 and ζ = 1 on Ω1 . Let 0 = min dist(spt(ζ), ∂Ω2 )/2. For all
0 <  < 0 , define u (x) = η ∗ u(x) for all x ∈ Ω2 , and set

v = η ∗ (ζ 2 u ,k ),k . (5.52)

Then v ∈ H01 (Ω) and can be used as a test function in (5.49); thus
Z Z
ij 2 
f η ∗ (ζ 2 u ,k ),k dx
   
a u,j η ∗ (ζ u ,k ),ki dx =
Ω Ω

or
Z Z
aij u,j η ∗(ζ 2 u ,i ),kk dx + aij u,j η ∗ (2ζζ,i u ),kk dx
   
Ω Z Ω Z
ij 
f η ∗ (ζ 2 u ,k ),k dx .
   
+ a u,j η ∗ (2ζζ,k u ),ki dx =
Ω Ω

The property of the convolution allows us to rewrite the equality above as


Z Z
η ∗ (aij u,j ) ,k (ζ 2 u ,i ),k dx = − f η ∗ (ζ 2 u ,k ),k dx
   
Ω Z Ω Z
ij 
aij u,j η ∗ (2ζζ,k u ),ki dx .
   
+ a u,j η ∗ (2ζζ,i u ),kk dx +
Ω Ω
 
Given the commutator ·, · defined in Definition 5.51,

η ∗ (aij u,j ) = η ∗, aij u,j + aij u ,j .


 
§5.9 Elliptic regularity 219

Therefore, we may further rewrite the previous integral equality as


Z Z
ζ a u ,jk u ,ik dx + 2 aij u ,jk ζζ,k u ,i dx
2 ij  
Ω Z Ω Z
ij  2 
η ∗, aij u,j ,k (ζ 2 u ,i ),k dx
  
+ a ,k u ,j (ζ u ,i ),k dx +
Ω Z Ω Z
2 
aij u,j η ∗ (2ζζ,i u ),kk dx
   
= − f η ∗ (ζ u ,k ),k dx +
ΩZ Ω

aij u,j η ∗ (2ζζ,k u ),ki dx .


 
+

By the commutation estimate (5.47),

η ∗, aij u,j ,k 2
  
L (Ω)
≤ CkakC 1 (Ω) kukH 1 (Ω) .

Therefore, by uniform ellipticity, the Cauchy-Young inequality, and the H 1 -


estimate (5.24), we find that
h
λ 0 kζD2 u k2L2 (Ω) ≤ C max kaij kL∞ (Ω) kukH 1 (Ω) kζD2 u kL2 (Ω)
i,j

+ kakC 1 (Ω) kukH 1 (Ω) kζD2 ukL2 (Ω)


i
+ kf kL2 (Ω) kζD2 ukL2 (Ω) + kukH 1 (Ω)


λ0 C
kζD2 u k2L2 (Ω) + kak2C 1 (Ω) + 1 kf k2L2 (Ω) ,


2 λ0
where we remark that C depends on the derivatives of ζ; thus C = C(Ω1 , Ω2 ).
Since ζ = 1 in Ω1 ,
C(Ω1 , Ω2 ) 
kD2 u kL2 (Ω1 ) ≤

kakC 1 (Ω) + 1 kf kL2 (Ω) .
λ0
Since the right-hand side does not depend on  > 0, there exists a subse-
quence
0
D2 u * w in L2 (Ω) .

By Theorem 2.78, u → u in H 1 (Ω1 ), so that w = D2 u on Ω1 . As weak


convergence is lower semi-continuous, kD2 ukL2 (Ω1 ) ≤ Ckf kL2 (Ω) . As Ω1 and
2 (Ω). For any w ∈ H 1 (Ω),
Ω2 are arbitrary, we have established that u ∈ Hloc 0
2 (Ω), we may integrate by parts to find
set v = ζw in (5.49). Since u ∈ Hloc
that Z
aij u,j ,i +f ζw dx = 0 ∀ w ∈ H01 (Ω) .
  

220 CHAPTER 5. Second Order Linear Elliptic Equations

Since w is arbitrary, and the spt(ζ) can be chosen arbitrarily close to ∂Ω, it
follows that for all x in the interior of Ω, we have that
∂  ij ∂u 
− a (x) (x) = f (x) for almost every x ∈ Ω . (5.53)
∂xi ∂xj
Step 2. Boundary estimates.
We proceed to establish the regularity of u all the way to the boundary
∂Ω. Let {Um }K
m=1 denote an open cover of Ω which intersects the boundary
∂Ω, and let {θm }K
m=1 denote a collection of charts such that

1. θm : B(0, rm ) → Um is a C ∞ -diffeomorphism,

2. det Dθm = 1,

3. θm (B(0, rm ) ∩ {xn = 0}) → Um ∩ ∂Ω,

4. θm (B(0, rm ) ∩ {xn > 0}) → Um ∩ Ω.

Let 0 ≤ ζm ≤ 1 in C0∞ (Um ) denote a partition of unity subordinate to the


open covering Um .
Substitute the test function
 −1
v = Λ [(ζm ◦ θm )2 Λ (u ◦ θm ),α ],α ◦ θm ∈ H01 (Ω)

into (5.49), and use the change of variables formula to obtain the identity
Z Z
ij ` k
a Aj (u◦θm ),` Ai (v ◦θm ),k dx = f ◦θm v ◦θm dx , (5.54)
B+ (0,rm ) B+ (0,rm )

where the C ∞ -matrix A(x) = [Dθm (x)]−1 and B+ (0, rm ) = B(0, rm )∩{xn >
0}. We define um = u ◦ θm , and bk` = aij A`j Aki . Then, with ξm = ζm ◦ θm ,
we can rewrite the test function as
 2
Λ um ,α ,α .

v ◦ θm = Λ ξm

Since differentiation commutes with convolution, we have that

2
(v ◦ θm ),k = Λ (ξm Λ um ,kα ),α +2Λ (ξm ξm ,k Λ um ,α ),α ,

and we can express the left-hand side of (5.54) as


Z
bk` (u ◦ θm ),` (v ◦ θm ),k dx = I1 + I2 ,
B+ (0,rm )
§5.9 Elliptic regularity 221

where
Z
I1 = bk` um ,` Λ (ξm
2
Λ um ,kα ),α dx ,
B+ (0,rm )
Z
I2 = 2 bk` um ,` Λ (ξm ξm ,k Λ um ,α ),α dx .
B+ (0,rm )

Next, we see that


Z
I1 = [Λ (bk` um ,` )],α (ξm
2
Λ um ,kα ) dx = I1a + I1b ,
B+ (0,rm )

where
Z
I1a = (bk` Λ um ,` ),α ξm
2
Λ um ,kα dx ,
B+ (0,rm )
Z
Λ , bk` um ,` ,α ξm
2
Λ um ,kα dx ,
  
I1b =
B+ (0,rm )

and where Λ , bk` is the commutator of the horizontal convolution operator


 

and multiplication. The integral I1a produces the positive sign-definite term
which will allow us to build the global regularity of U , as well as an error
term:
Z h i
2 k`
I1a = ξm b Λ um ,`α Λ um ,kα + bk` ,α Λ um ,` ξm
2
Λ um ,kα dx ;
B+ (0,rm )

thus, together with the right hand-side of (5.54), we see that


Z Z

2 k` m m k` m 2 m
ξm b Λ u ,`α Λ u ,kα dx ≤ b ,α Λ u ,` ξm Λ u ,kα dx

B+ (0,rm ) B+ (0,rm )
Z

+ |I1b | + |I2 | + f ◦ θm v ◦ θm dx .

B+ (0,rm )

Since each θm is a C ∞ -diffeomorphism, it follows that the matrix bk` is


positive definite: there exists λ > 0 such that

λ|Y |2 ≤ bk` Yk Y` ∀Y ∈ Rn .

It follows that
Z Z

2 ¯
λ ξm |∂DΛ um |2 dx ≤ bk` ,α Λ um ,` ξm
2
Λ um ,kα dx

B+ (0,rm ) B+ (0,rm )
Z

+ |I1b | + |I2 | + f ◦ θm v ◦ θm dx ,

B+ (0,rm )
222 CHAPTER 5. Second Order Linear Elliptic Equations

where D = (∂x1 , ..., ∂xn ) and ∂¯ = (∂x1 , ..., ∂xn−1 ). Application of the Cauchy-
Young inequality with δ > 0 shows that
Z Z

bk` ,α Λ um ,` ξm2
Λ um ,kα ] dx + |I2 | + f ◦ θm v ◦ θm dx


B+ (0,rm ) B+ (0,rm )
Z
2 ¯
≤δ ξm |∂DΛ um |2 dx + Cδ kf k2L2 (Ω) .
B+ (0,rm )

It remains to establish such an upper bound for |I1b |. Nevertheless, by the


commutator estimate (5.46),
 
∂xα Λ , bk` um ,` 2

L (B+ (0,rm ))
≤ CkbkW 1,∞ ((B+ (0,rm )) kukH 1 (Ω) ≤ CkakC 1 (Ω) kf kL2 (Ω) .

It then follows from the Cauchy-Young inequality with δ > 0 that


Z
2 ¯
|I1b | ≤ δ ξm |∂DΛ um |2 dx + Cδ kf |2L2 (Ω) .
B+ (0,rm )

By choosing 2δ < λ, we obtain the estimate


Z
2 ¯
ξm |∂DΛ um |2 dx ≤ Cδ kf k2L2 (Ω) .
B+ (0,rm )

Since the right hand-side is independent of , we find that


Z
2 ¯
ξm |∂Dum |2 dx ≤ Cδ kf k2L2 (Ω) . (5.55)
B+ (0,rm )

From (5.53), we know that − aij (x)u,j (x) ,i = f (x) for a.e. x ∈ Um .


By the chain-rule this means that almost everywhere in B+ (0, rm ),

−bk` um ,k` = bk` ,k um ,` +f ◦ θm ,

or equivalently,

−bnn um kα m β` m k` m
nn = b u ,αk +b u ,`β +b ,k u ,` +f ◦ θm .

If en = (0, · · · , 0, 1), then bnn = eT 2


n b en ≥ λ|en | = λ > 0; thus it follows
from (5.55) that
Z
2
ξm |D2 um |2 dx ≤ Cδ kf k2L2 (Ω) . (5.56)
B+ (0,rm )

Summing over m from 1 to K and combining with our interior estimates,


we have that

kukH 2 (Ω) ≤ Ckf kL2 (Ω) . 


§5.9 Elliptic regularity 223

5.9.2 Dirichlet problem with inhomogeneous boundary con-


dition

Next we want to solve the following Dirichlet problem with inhomogeneous


boundary condition:

∂  ij ∂u 
− a =f in Ω, (5.57a)
∂xi ∂xj
u=g on ∂Ω , (5.57b)

where f ∈ L2 (Ω) and g ∈ H 1.5 (∂Ω). By Theorem 2.156, there exists φ ∈


H 2 (Ω) so that φ = g on ∂Ω and kφkH 2 (Ω) ≤ CkgkH 1.5 (∂Ω) . Then w = u − φ
is the solution to

∂  ij ∂w  ∂  ij ∂φ 
− a =f+ a in Ω , (5.58a)
∂xi ∂xj ∂xi ∂xj
w=0 on ∂Ω , (5.58b)

hence u satisfies the estimate


h i h i
kuk2H 2 (Ω) ≤ C kwk2H 2 (Ω) + kφk2H 2 (Ω) ≤ C kf k2L2 (Ω) + kgk2H 1.5 (∂Ω) . (5.59)

5.9.3 Neumann problem

In this subsection we consider the Neumann problem

∂  ij ∂u 
− a =f in Ω, (5.60a)
∂xi ∂xj
∂u
aij Ni = g on ∂Ω , (5.60b)
∂xj

Note that the weak formulation for the Neumann problem is

ij ∂u ∂ϕ
Z Z Z
a dx = f v dx + gv dS ∀ v ∈ H 1 (Ω) .
Ω ∂x j ∂xi Ω ∂Ω

In order to obtain the corresponding


Z regularity result, it suffices to estimate
the additional boundary integral gv dS for test functions
∂Ω

1. v = η ∗ (ζ 2 u ,k ),k in the proof of the interior estimates, or


 −1
2. v = Λ [(ζm ◦ θm )2 Λ (u ◦ θm ),α ],α ◦ θm in the proof of the boundary
estimates.
224 CHAPTER 5. Second Order Linear Elliptic Equations

There is nothing we have to do if v has compact support in Ω, so we


consider the case that
 −1
v = Λ [(ζm ◦ θm )2 Λ (u ◦ θm ),α ],α ◦ θm

for some fixed m = 1, 2, · · · , K. However, it follows from H 0.5 (∂Ω)-H −0,5 (∂Ω)
duality the Trace Theorem, and the Cauchy-Young inequality that
Z
gϕ dS ≤ CkgkH 0.5 (∂Ω) ζm Λ um ,α H 0.5 (∂Ω∩Um )

∂Ω
≤ Cδ kgk2H 0.5 (∂Ω) + δ ζm ∂xα Λ um H 1 (Um )

h i 2
≤ Cδ kf k2L2 (Ω) + kgk2H 0.5 (∂Ω) + δ ζm ∂xα DΛ um L2 (Um ) .

By choosing δ > 0 small enough, we obtain


h i
kukH 2 (Ω) ≤ C kf kL2 (Ω) + kgkH 0.5 (∂Ω) . (5.61)

5.9.4 The presence of b and c in the elliptic operator L

The H 2 -regularity of the weak solution to the Dirichlet problem with bi =


c = 0 for all i is studied in Theorem 5.54. With the inclusion of b and c,
the existence of a solution is non-trivial; nevertheless, can be verified by
the Fredholm alternative. As long as the existence of a weak solution is
guaranteed, we are interested in the regularity of weak solutions. We have
the following

Theorem 5.55. Assume that

aij , bi , c ∈ C 1 (Ω) , f ∈ L2 (Ω) .

Suppose that u ∈ H01 (Ω) is a weak solution of the Dirichlet problem (5.41),
and ∂Ω is of class C 2 . Then u ∈ H 2 (Ω) ∩ H01 (Ω) and satisfies the estimate
h i
kukH 2 (Ω) ≤ C kf kL2 (Ω) + kukL2 (Ω) (5.62)

for some constant C depending only on m, U , and the coefficient of L.

Proof. If u is a weak solution to (5.41), then


Z Z
aij u,j ϕ,i dx = F ϕ dx ∀ ϕ ∈ H01 (Ω) ,
Ω Ω
§5.9 Elliptic regularity 225

where F = f − b · Du − cu . Note that this is the same as saying that u is


the unique weak solution to the Dirichlet problem
∂  ij ∂u 
− a =F in Ω ,
∂xi ∂xj
u=0 on ∂Ω .

Therefore, by Theorem 5.54,

kukH 2 (Ω) ≤ CkF kL2 (Ω) ,

and (5.62) follows from Theorem 5.50 that


h i h i
kF kL2 (Ω) ≤ C kf kL2 (Ω) + kukH 1 (Ω) ≤ C kf kL2 (Ω) + kukL2 (Ω) . 

5.9.5 Higher regularity

The proof to the following two theorems, depending on choices of different


test functions, are similar to the proof of Theorem 5.54 so we state without
proving the theorems.

Theorem 5.56. Let m be a non-negative integer, and assume that

aij , bi , c ∈ C m+1 (Ω) , f ∈ H m (Ω) .

Suppose that u ∈ H01 (Ω) is a weak solution of the Dirichlet problem (5.41),
and ∂Ω is of class C m+2 . Then u ∈ H m+2 (Ω) ∩ H01 (Ω) and satisfies the
estimate
h i
kukH m+2 (Ω) ≤ C kf kH m (Ω) + kukL2 (Ω)

for some constant C depending only on m, Ω, and the coefficient of L.

Theorem 5.57. Let m be a non-negative integer, and assume that

aij , bi , c ∈ C m+1 (Ω) , f ∈ H m (Ω) .

Suppose that u ∈ H 1 (Ω) is a weak solution of the boundary value problem


∂  ij ∂u  ∂u
− a + bi + cu = f in Ω , (5.63a)
∂xi ∂xj ∂xi
∂u
aij Ni = g on ∂Ω , (5.63b)
∂xj
226 CHAPTER 5. Second Order Linear Elliptic Equations

and ∂Ω is of class C m+2 . Then u ∈ H m+2 (Ω) and satisfies the estimate
h i
kukH m+2 (Ω) ≤ C kf kH m (Ω) + kgkH m+0.5 (∂Ω) + kukL2 (Ω)

for some constant C depending only on m, Ω, and the coefficient of L.

5.10 Two-phase problems - Elliptic equations with


mixed type of boundary condition

In continuum mechanics, it is quite natural to consider multi-phase prob-


lems. For example, two immiscible fluids, such as water and oil, present
in the same fluid container. While the dynamics of fluids inside different
phases looks independent, they interact through the common interface thus
the behavior of the fluid could be quite complicated.
In the following, we propose a two-phase elliptic problem motivated by
the two-phase Hele-Shaw equation, and study the existence, uniqueness,
and regularity of the solution. Suppose that Ω1 and Ω2 are two disjoint
connected domain separated by a common boundary Γ, called the interface
between Ω1 and Ω2 , so that Ω = Ω1 ∪ Ω2 is connected (thus ∂Ω2 = ∂Ω ∪ Γ).
We call Ω1 and Ω2 the outer phase and the inner phase, respectively. We
consider

−div(b1 Dp1 ) = f1 in Ω1 , (5.64a)


−div(b2 Dp2 ) = f2 in Ω2 , (5.64b)
p1 = p2 on Γ, (5.64c)
(b1 Dp1 − b2 Dp2 ) · N = 0 on Γ, (5.64d)
p1 = 0 on ∂Ω , (5.64e)

where N on Γ is chosen to be a fixed orientation, and b1 , b2 are positive


functions with positive lower bounds, that is, b1 ≥ λ 0 and b2 ≥ λ 0 for some
constant λ 0 > 0.
Equation (5.64) might look over-constrained because it looks like two
separated elliptic equations with both Dirichlet and Neumann boundary
conditions. However, the following example indicates that (5.64) will not be
well-posed if imposing only (5.64c) or only (5.64d) as the interface condition.
§5.10 Two-phase problems 227

Example 5.58. Let Ω1 = (−2, −1) ∪ (1, 2) and Ω2 = (−1, 1). Then for all
a, b ∈ R, p1 and p2 given by

p1 (x) = x2 + ax + (2a − 4) 1(−2,−1) (x) + x2 + (a − 4)x + (4 − 2a) 1(1,2) (x) ,


   

p2 (x) = (a − 2)x + b ,

are solutions to

−p001 (x) = −2 for all x ∈ Ω1 ,


−p002 (x) = 0 for all x ∈ Ω2 ,
p10 (x) − p20 (x) = 0 at x = −1 and x = 1 ,
p1 (x) = 0 at x = −2 and x = 2 .

Similarly, for all a, c ∈ R , p1 and p2 given by

p1 (x) = x2 + ax + (2a − 4) 1(−2,−1) (x) + x2 + cx − (2c + 4) 1(1,2) (x) ,


   

a+c a−c−6
p2 (x) = − x+ ,
2 2
are solutions to

−p001 (x) = −2 for all x ∈ Ω1 ,


−p002 (x) = 0 for all x ∈ Ω2 ,
p1 (x) = p2 (x) at x = −1 and x = 1 ,
p1 (x) = 0 at x = −2 and x = 2 .

However, this is only one solution,

p1 (x) = x2 + 2x 1(−2,−1) (x) + x2 − 2x 1(1,2) (x) ,


   

p2 (x) = −1 ,

to the equation

−p001 (x) = −2 for all x ∈ Ω1 ,


−p002 (x) = 0 for all x ∈ Ω2 ,
p1 (x) = p2 (x) at x = −1 and x = 1 ,
p10 (x) + p20 (x) = 0 at x = −1 and x = 1 ,
p1 (x) = 0 at x = −2 and x = 2 .
228 CHAPTER 5. Second Order Linear Elliptic Equations

5.10.1 The variational formulation and weak solutions

Suppose for now that there exists smooth solution p1 and p2 to the problem.
Testing (5.64a) against a test function ϕ ∈ H 1 (Ω2 ) and integrating by parts,
we find that
∂p1
Z Z Z
b1 Dp1 · Dϕ dx − b1 (τ1 ϕ) dS = f1 ϕ dx , (5.65)
Ω1 Γ ∂N1 Ω1

where N1 is the outward-pointing unit normal of Ω1 , and τ1 is the trace


operator on Ω1 . Similarly, testing (5.64b) against a test function ϕ ∈ H 1 (Ω1 )
so that ϕ = 0 on ∂Ω (ϕ might not vanish on Γ) and integrate by parts, we
find that
∂p2
Z Z Z
b2 Dp2 · Dϕ dx − b2 (τ2 ϕ) dS = f2 ϕ dx (5.66)
Ω2 Γ ∂N2 Ω1

where N2 is the outward-pointing unit normal of Ω2 , andτ2 is the trace


operator on Ω2 . Since N1 = −N2 on Γ,
∂p1 ∂p1
Z Z
b1 (τ1 ϕ) dS + b2 (τ1 ϕ) dS = 0 if τ1 ϕ = τ2 ϕ .
Γ ∂N1 Γ ∂N1
Therefore, for all ϕ ∈ H01 (Ω) , summing (5.65) and (5.65) together we obtain
that for all ϕ ∈ H01 (Ω),
Z Z Z Z
b1 Dp1 · Dϕ dx + b2 Dp2 · Dϕ dx = f1 ϕ dx + f2 ϕ dx .
Ω1 Ω2 Ω1 Ω2

Let p = p1 1Ω1 + p2 1Ω2 , b = b1 1Ω1 + b2 1Ω2 , f = f1 1Ω1 + f2 1Ω2 , then the


integral identity above implies that (5.64) can be rewritten as

−div(bDp) = f in Ω, (5.67a)
p=0 on ∂Ω . (5.67b)

The discussion above leads to the following

Definition 5.59. For any f ∈ H −1 (Ω), the variational form of (5.64) is


Z Z
b1 Dp1 · Dϕ dx + b2 Dp2 · Dϕ dx = hf, ϕiH01 (Ω) ∀ ϕ ∈ H01 (Ω) .
Ω1 Ω2

(p1 , p2 ) is called a weak solution to (5.64) if it satisfies the variational for-


mulation.

Since (5.64) can be rewritten as a one-phase Dirichlet problem (5.67),


by Theorem 5.38 there exists a unique weak solution p to the problem.
5.11. MAXIMUM PRINCIPLE 229

5.10.2 Regularity

In elliptic regularity, the coefficient aij are assumed to be C k (Ω) in order


to obtain u ∈ H k+1 (Ω). However, in this two-phase problem, b1 does not
have to be identical to b2 on the interface Γ. In other words, the coefficient
b in (5.67) might be discontinuous in Ω; thus the coefficient b will possess
limited regularity, say b ∈ L∞ (Ω). However, b1 and b2 can still be smooth
in Ω1 and Ω2 . This is the case that we are going to explore, and we have
the following

Theorem 5.60. Suppose that b1 ∈ C 1 (Ω1 ) and b2 ∈ C 1 (Ω2 ) have positive


lower bounds. Then for any f ∈ L2 (Ω) , the weak solution p ∈ H01 (Ω) to
(5.64) in fact belongs to the space

H 2 (Ω1 ) ∩ H 2 (Ω2 ) ≡ p ∈ H01 (Ω) p ∈ H 2 (Ω1 ) and p ∈ H 2 (Ω2 ) ,




and satisfies

kpkH 2 (Ω1 ) + kpkH 2 (Ω2 ) ≤ Ckf kL2 (Ω) .

5.11 Maximum principle

To study the maximum principle, we assume that the elliptic operator L is


of non-divergence form, that is,

∂2u ∂u
Lu = −aij + bi + cu .
∂xi ∂xj ∂xi

We remark here that L in divergence form can be rewritten as the non-


divergence form above if aij ∈ C 1 (Ω).
The following results generalize the maximum principle in Chapter 4.

Theorem 5.61 (Weak maximum principle). Assume u ∈ C 2 (Ω) ∩ C 0 (Ω)


and c ≥ 0 in Ω. Let u+ = max{u, 0} and u− = max{−u, 0}.

(i) If Lu ≤ 0 in Ω, then max u ≤ max u+ .


Ω ∂Ω

(ii) If Lu ≥ 0 in Ω, then min u ≤ − max u− .


Ω ∂Ω

Proof. Will be added later. 


230 CHAPTER 5. Second Order Linear Elliptic Equations

Lemma 5.62 (Hopf’s lemma). Assume u ∈ C 2 (Ω) ∩ C(Ω) and c ≡ 0 in Ω.


Suppose further that Lu ≤ 0 in Ω, and x0 ∈ ∂Ω is a point with the property
that

u(x0 ) > u(x) ∀x ∈ Ω,

and there exists an open ball B ⊆ Ω with x0 ∈ ∂B (this is call the interior
∂u
ball condition at x0 ). Then (x0 ) > 0, where N is the outward pointing
∂N
unit normal to Ω an x0 .

Proof. Assume that B = B(y, r) for some y ∈ Ω and r > 0. Define


2 2
v(x) = e−λ|x−y| − e−λr , x ∈ B(y, r)

for some λ > 0 to be chosen later. Then if A = [aij ] and b = (b1 , · · · , bn ),

∂2v ∂v
Lv = −aij + bi
∂xi ∂xj ∂xi
2
h  i
= e−λ|x−y| aij − 4λ2 (xi − yi )(xj − yj ) + 2λδ ij − 2λbi (xi − yi )


2
≤ e−λ|x−y| − 4θλ2 |x − y|2 + 2λtr(A) + 2λ|b||x − y| .


By defining the open annular region R = B(y, r) − B(y, r/2), we have


2
Lv ≤ e−λ|x−y| − 4θλ2 |x − y|2 + 2λtr(A) + 2λ|b||x − y| ≤ 0
 
in R

if λ > 0 is large enough.


Now let w = u + v − u(x0 ). Then w ≤ 0 on ∂R, and Lw ≤ 0 in R; hence
the weak maximum principle (Theorem 5.61) implies that w ≤ 0 in R. On
the other hand, w(x0 ) = 0; thus
∂w ∂u ∂v
0≤ (x0 ) = (x0 ) +  (x0 )
∂N ∂N ∂N
which implies
∂u ∂v |x0 − y|2 −λ|x0 −y|2 2
(x0 ) ≥ − (x0 ) = 2λ e = 2λre−λr > 0. 
∂N ∂N r
Theorem 5.63 (Strong maximum principle). Suppose that Ω is a connected,
open, bounded domain of Rn , and u ∈ C 2 (Ω) ∩ C(Ω) and c ≡ 0 in Ω.

(i) If Lu ≤ 0 in Ω, and u attains its maximum over Ω at an interior point


x ∈ Ω, then u is constant within Ω.
5.12. EIGENVALUES AND EIGENFUNCTIONS 231

(ii) If Lu ≥ 0 in Ω, and u attains its minimum over Ω at an interior point


x ∈ Ω, then u is constant within Ω.

Proof. Let M = max u, A = {x ∈ Ω | u(x) = M }, and S = {x ∈ Ω | u(x) <



M }. If u 6≡ M , then S 6= ∅ and S is an open set . Choose a point y ∈ S such
that dist(y, A) < dist(y, ∂Ω), and let B denote the largest ball with center
y whose interior lies in S. Then for some x0 interior point of A, x0 ∈ ∂B.
∂u
Therefore, by the Hopf lemma, (x0 ) > 0, but this contradicts to the fact
∂N
that if Du(x0 ) = 0. 

5.12 Eigenvalues and eigenfunctions

In this section, we consider the following eigenvalue problem

Lw = λw in Ω, (5.68a)
w=0 on ∂Ω , (5.68b)

where Ω is an open bounded smooth domain, and λ is called an eigenvalue


of L if there exists non-trivial w satisfying (5.68), in which case w is called
an eigenfunction of L.
For simplicity, we consider L with divergence form, and b = c = 0, that
is,
∂  ij ∂u 
Lu = − a .
∂xi ∂xj

We assume that a ∈ C ∞ (Ω), aij = aji , and L is strictly elliptic in Ω.

Definition 5.64 (Spectrum). The collection of eigenvalues of L is called


the spectrum of L.

Theorem 5.65. Let Σ be the spectrum of L. Then Σ ⊆ R, and Σ is at most


countable. Let Σ = {λk }∞
k=1 , where the eigenvalue is counted according to
its multiplicity, then

0 < λ1 ≤ λ2 ≤ · · ·

and λk → ∞ as k → ∞. Finally, there exists an orthonormal basis {wk }∞


k=1
of L2 (Ω), where wk ∈ H01 (Ω) is an eigenfunction corresponding to λk , that
232 CHAPTER 5. Second Order Linear Elliptic Equations

is,

Lwk = λk wk in Ω,
wk = 0 on ∂Ω .

Remark 5.66. By the elliptic regularity Theorem 5.56, wk ∈ C ∞ (Ω) for all
k ∈ N.

Proof. Let S = L−1 . Then S : L2 (Ω) → L2 (Ω) is a bounded, linear, compact


operator. If λ is an eigenvalue of L, then λ−1 is an eigenvalue of S. By
Theorem B.31, we know that σ(S), the spectrum of S, is at most countable,
and has 0 as the unique limit point.
We claim that S is symmetric; that is, for f, g ∈ L2 (Ω), we have that
(Sf, g)L2 (Ω) = (f, Sg)L2 (Ω) . With u = Sf and v = Sg,

Lu = f in Ω, Lv = g in Ω,
and
u=0 on ∂Ω , v=0 on ∂Ω ,
so that
∂v ∂u
Z
(Sf, g)L2 (Ω) = (u, Lv)L2 (Ω) = aij dx
Ω ∂xj ∂xi
∂u ∂v
Z
= aij dx = (Lu, v)L2 (Ω) = (f, Sg)L2 (Ω) .
Ω ∂xj ∂xi
The symmetry of S implies that Σ ⊆ R. The unique solvability of L for λ < 0
implies that Σ ⊆ R+ . Since σ(S) has 0 as the unique limit point, λk → ∞.
By Theorem B.38, the eigenfunctions of S make up an orthonormal basis
of L2 (Ω), and eigenfunctions of S are eigenfunctions of L as long as the
corresponding eigenvalues are not zero. 
n o∞
Corollary 5.67. √wλk is an orthonormal basis of H01 (Ω), here the
k k=1
inner product of H01 (Ω) is defined by (u, v)H01 (Ω) = (Du, Dv)L2 (Ω) .

Proof. Since (wk , w` )H01 (Ω) = λk δk` , it suffices to show that if u ∈ H01 (Ω) so
that (u, wk )H01 (Ω) = 0 for all k ∈ N, then u = 0. However,

(u, wk )H01 (Ω) = λk (u, wk )L2 (Ω) ,

so (u, wk )H01 (Ω) = 0 implies that u is orthogonal to wk . The completeness of


{wk }∞
k=1 implies that u = 0. 
§5.12 Eigenvalues and eigenfunctions 233

Definition 5.68. We call λ1 > 0 the principal eigenvalue of L.

Theorem 5.69. The principal eigenvalue λ1 is simple, and the correspond-


ing eigenfunction w1 has only one sign, that is, w1 > 0 in Ω or w1 < 0 in
Ω. Moreover,
nZ ∂u ∂u o
λ1 = min aij dx u ∈ H01 (Ω) , kukL2 (Ω) = 1 . (5.69)

Ω ∂xj ∂xi

Proof. We first prove (5.69). Since {wk }∞


k=1 is an orthonormal basis of
L2 (Ω), for a given u ∈ L2 (Ω) with kukL2 (Ω) = 1,

X
u= dk wk , (5.70)
k=1


d2k = kuk2L2 (Ω) = 1.
P
where dk = (u, wk )L2 (Ω) . By the Parseval identity,
k=1
ij ∂u ∂v
Z
Let B[u, v] ≡ a ij
dx. Since a is symmetric, B(·, ·) can be
Ω ∂xj ∂xi
viewed as an inner product of H01 (Ω), and { √wλk }∞k=1 forms an orthonormal
k
basis of H01 (Ω) with respect to this new inner product (because B(wk , wk ) =
λk kwk k2L2 (Ω) = λk and B(wk , w` ) = λk (wk , w` )L2 (Ω) = 0 if k 6= `). Moreover,

X wk
u= µk √
k=1
λk

with µk = B(u, √wλk ). By (5.70) µk = dk λk ; hence
k


X ∞
X
B[u, u] = µ2k = d2k λk ≥ λ1 .
k=1 k=1

Since B[w1 , w1 ] = λ1 , (5.69) is valid.


Claim: If u ∈ H01 (Ω) and kukL2 (Ω) = 1, then u is a weak solution of the
eigenvalue problem

Lu = λ1 u in Ω, (5.71a)
u=0 on ∂Ω , (5.71b)

if and only if B[u, u] = λ1 .


Proof of claim: The sufficiency is obvious. To prove the necessity, suppose
that u ∈ H01 (Ω), kukL2 (Ω) = 1 satisfies B[u, u] = λ1 . Let dk = (u, wk )L2 (Ω) ,
234 CHAPTER 5. Second Order Linear Elliptic Equations

then

X ∞
X ∞
X
λ1 d2k = λ1 = B[u, u] = d2k λk ⇒ d2k (λk − λ1 ) = 0 .
k=1 k=1 k=1

As a consequence, dk = 0 if λk > λ1 . Since λ1 has finite multiplicity,


m
X
u= (u, wk )wk
k=1

for some m ∈ N, where Lwk = λ1 wk for k = 1, · · · , m. But then Lu = λ1 u


which concludes the necessity. End of proof of claim.

Now let u ∈ H01 (Ω) with kukL2 (Ω) = 1 be a non-trivial weak solution of
(5.71). Then u+ and u− satisfy
( (
+ Du a.e. on {u ≥ 0} − 0 a.e. on {u ≥ 0}
Du = Du =
0 a.e. on {u ≤ 0} , −Du a.e. on {u ≤ 0} .

Moreover, B[u+ , u− ] = 0. Therefore,

λ1 = B[u, u] = B[u+ , u+ ] + B[u− , u− ]


by (5.69)
≥ λ1 ku+ k2L2 (Ω) + λ1 ku− k2L2 (Ω) = λ1 ;

hence the inequality above is in fact an equality. As a consequence,

B[u+ , u+ ] = λ1 ku+ k2L2 (Ω) and B[u− , u− ] = λ1 ku− k2L2 (Ω) ;

thus u+ and u− are both weak (in fact smooth) solutions to (5.71). In
particular, Lu+ ≥ 0 in Ω, and the strong maximum principle (Theorem
5.63) implies that u+ > 0 in Ω or u+ ≡ 0 in Ω. Same arguments apply to
u− , so u has only one sign in Ω.
Finally suppose
Z that u1 and u2 are two non-trivial weak solutions to
(5.71). Then uk dx 6= 0 for k = 1, 2. Therefore, there exists a constant

χ 6= 0 such that
Z
(u1 − χu2 )dx = 0 .

Since u1 −χu2 is again a weak solution to (5.71), it follows that u1 −χu2 ≡ 0


in Ω, hence the eigenvalue λ1 is simple. 
5.13. EXERCISES 235

5.13 Exercises

Problem 5.1. The Lax-Milgram theorem cannot directly be applied to the


Robin problem (5.29) in the case that k is negative. This, however, does not
imply that there does not exist a solution to the Robin problem with k < 0.
This exercise is concerned with the solvability of the Robin problem in
du
the one-dimensional setting. For Ω = (a, b), we introduce the notation
dn
to mean
(
du a + b 0 u0 (x) if x = b ,
(x) = sgn x − u (x) =
dn 2 −u0 (x) if x = a .

Complete the following.

1. Find the unique solution to the Robin problem

−u00 (x) = −2 for all x ∈ (0, 1) ,


du
(x) − u(x) = 0 at x = 0 and x = 1 .
dn

2. Find the non-trivial solutions to the Robin problem

−u00 (x) = 0 for all x ∈ (0, 2) ,


du
(x) − u(x) = 0 at x = 0 and x = 2 .
dn

3. Show that there is no solution to the Robin problem

−u00 (x) = 6x for all x ∈ (0, 2) ,


du
(x) − u(x) = 0 at x = 0 and x = 2 .
dn

4. Next, suppose that Ω = x ∈ R2 |x| < 1 . Prove that for all f ∈


L2 (Ω) there exists a unique weak solution to

u − ∆u = f in Ω,
du
− ∆g u = 0 on ∂Ω .
dn

Problem 5.2. Let η = (η1 , η2 ) be a vector-valued function such that

(a) η1 : [0, 1] → [0, 1] is continuous, and η1 (0) = 0 and η1 (1) = 1;


236 CHAPTER 5. Second Order Linear Elliptic Equations

(b) η1 : (0, 1) → (0, 1) is a diffeomorphism;

(c) η2 : [0, 1] → R is continuous;

(d) η2 : (0, 1) → R is differentiable.

Define g = η 0 · η 0 = |η 0 |2 , n = g −1/2 (−η20 , η10 ), and


0
H(η) = −g −1/2 g −1/2 η 0 · n .

Complete the following.

1. Show that H(η) = −g −1 (η 00 · n).

2. Let η  be a family of functions satisfying (a)-(d) with η 0 (x) = (x, 0)


d
and η  = u. Show that
d =0
d
L1 u ≡ H(η  ) = −u002 .
d =0
This quantity is called the linearization of H(η) about η = η 0 .

3. Find a solution to L1 u = 1. Is this solution unique?

4. Let θ : (0, 1) → (0, 1) be a diffeomorphism. Show that

H(η ◦ θ) = H(η) ◦ θ .

Therefore, if η is a solution to H(η) = 1 , then η ◦ θ is also a solution


for every diffeomorphism θ. Consequently, the kernel of H is infinitely
dimensional.

5. Since η1 : (0, 1) → (0, 1) is a diffeomorphism, η1 has a unique inverse


η1−1 . Define h = η2 ◦ η1−1 . Then η ◦ η1−1 (x) = x, h(x) . Show that


−3/2 0
H(h) ≡ H(η ◦ η1−1 ) = 1 + |h0 |2 h00 = (1 + |h0 |2 )−1/2 h0 .

6. Let h0 be a given function on (0, 1) satisfying (c) and (d). Find the
linearization of H(h) about h = h0 . In other words, compute
d
L2 δh ≡ H(h )
d =0
d

given that h0 = h0 and h = δh .
d =0
§5.13 Exercise 237

7. Show that the solution to L2 δh = 1 is unique for any give h0 .

Problem 5.3. Let Ω ⊆ Rn denote an open, bounded, and smooth domain.


Let u : Ω → Rn denote a vector-valued function, and p : Ω → R be a scalar
function.

1. Find a unique solution to the PDE

u + Dp = 0 in Ω, (5.72a)
divu = 0 in Ω, (5.72b)
u·N=g on ∂Ω . (5.72c)

2. Find a unique solution to (5.72a,b) with (5.72c) replaced by the bound-


ary condition

p=g on ∂Ω .

3. Find an estimate for u and p in terms of the given data g for parts 1.
and 2.

4. Next, suppose that f ∈ H 1 (Ω) and that

u + Dp = f in Ω,
divu = 0 in Ω,
p = −∆g (u · N) on ∂Ω ,

where N denotes the outward unit normal to ∂Ω. Prove that there
exists a unique solution to this problem and find the estimates verified
by u and p.

Problem 5.4. Recall that

1
(Ω; R3 ) ≡ w ∈ H 1 (Ω; R3 ) divw = 0, w · N = 0 on ∂Ω .

Hdiv

We define the Leray projector P : H 1 (Ω; R3 ) → Hdiv


1 (Ω; R3 ) as follows: for

a vector u ∈ H 1 (Ω; R3 ), find w ∈ Hdiv


1 (Ω; R3 ) and p ∈ H 2 (Ω) such that

u = w + Dp. As long as the decomposition is unique, Pu ≡ w.

1. What is the (elliptic) equation that p must satisfy?


238 CHAPTER 5. Second Order Linear Elliptic Equations

2. Is the decomposition u = w + Dp unique? In other words, is it possible


to find (w1 , p1 ) 6= (w2 , p2 ) such that w1 + Dp1 = w2 + Dp2 ?

3. Show that P : H 1 (Ω; R3 ) → Hdiv


1 (Ω; R3 ) is a continuous projection,

that is,

kPukH 1 3 ≤ CkukH 1 (Ω;R3 ) and P(Pu) = u ∀ u ∈ H 1 (Ω; R3 ) .


div (Ω;R )

Problem 5.5. In this problem, we examine if the specification of the diver-


gence and curl of a vector with periodic boundary conditions is sufficient to
uniquely determine the vector field.
Let u : T3 → R3 be a vector-valued function defined on T3 .

1. Show that curlcurlu = −∆u + Ddivu for all u ∈ C 2 (T3 ).

2. Show that

kDuk2L2 (T3 ) = kdivuk2L2 (T3 ) + kcurluk2L2 (T3 ) ∀ u ∈ H 1 (T3 ) , (5.74)

3
X
where kDuk2L2 (T3 ) ≡ kui ,j k2L2 (T3 ) .
i,j=1

3. Given f ∈ H −1 (T3 ) and g ∈ H −1 (T3 ; R3 ), find a vector field u ∈


L2 (T3 ; R3 ) with divu = f and curlu = g. Is this u unique? (Hint:
Write u in the form u = Dp + curlw for some p ∈ H 1 (T3 ) and w ∈
1 (T3 ; R3 ).)
Hdiv

Problem 5.6 (Calderon-Zygmund-type inequality). Let u ∈ H 3 (R3 ; R3 )


with divu = 0 and let ω = curlu. Then, with log+ |f | := log max(1, |f |),
h i
kDukL∞ (R3 ) ≤ C 1 + 1 + log+ kukH 3 (R3 ) kωkL∞ (R3 ) + kωkL2 (R3 ) .


This problem extends the basic elliptic estimate


h i
kukW 1,p (R3 ) ≤ Cp kcurlukLp (R3 ) + kdivukLp (R3 ) 1 < p < ∞,

which reads for the above u as

kukW 1,p (R3 ) ≤ Cp kωkLp (R3 ) 1 < p < ∞,


§5.13 Exercise 239

where the constant Cp → ∞ as p → ∞. The lemma shows that in the limit


as p → ∞, a logarithmic correction enters the estimate.
The proof of this result relies on the so-called Biot-Savart kernel which
is fundamental in fluid dynamics and electromagnetism. Since divu = 0, we
may introduce a vector potential ψ so that

u = −curlψ .

Then ω = curlu = ∆ψ and this implies that ψ = ∆−1 ω; thus, using the
fundamental solution for Poisson’s equation on R3 ,
1 ω(y)
Z
ψ(x) = − dy ,
4π R3 |x − y|
and hence
1 x−y
Z
u(x) = −curlψ(x) = − × ω(y)dy .
4π R3 |x − y|3
x
Definition 5.70. K(x)f = × f is called the Biot-Savart kernel, so
4π|x|3
that
Z Z
u(x) = K(x − y) ω(y)dy = K(y) ω(x − y)dy .
R3 R3
Use this definition to establish the result.
Hint. In computing Du(x), you must differentiate under the integral sign,
a process that leads to a singular integral that must be interpreted as a p.v.
integral, i.e.,
Z Z
K(y)Dω(x − y)dy := lim K(y)Dω(x − y)dy .
R3 δ→0 |y|≥δ

Problem 5.7. Show that for any p ∈ [1, n), there exist constants α ∈ (0, 1),
ν ∈ (0, 1], and a real symmetric n × n matrix a(x) = aij (x) satisfying the
 

uniform ellipticity condition

ν|ξ|2 ≤ aij (x)ξi ξj ≤ ν −1 |ξ|2 for all ξ ∈ Rn and x ∈ Rn ,

and such that the function u(x) ≡ 1 − |x|α satisfies

Lu ≡ aij u,ij = 0 a.e. in B(0, 1) ⊆ Rn ,

and its derivatives D2 u ∈ Lp (B(0, 1)). This problem shows that the maxi-
mum principle does not hold in general for uniformly elliptic equation with
discontinuous coefficients in the Sobolev class W 2,p , when p < n.
240 CHAPTER 5. Second Order Linear Elliptic Equations

Problem 5.8. Let ψ : U ⊆ Rn → V ⊆ Rn be a diffeomorphism, and


f : V → R is a smooth function. Let x ∈ U and y = ψ(x) ∈ V be one
coordinate system on U and V. Show that
∂f ∂(f ◦ ψ)
◦ ψ = Aji ,
∂yi ∂xj

where A = (Dψ)−1 .
Hint: Recall the chain rule in the theory of partial derivatives of a function
of multiple variables.

Problem 5.9. Given f ∈ L2 (Ω). Let Kf ∈ H 2 (Ω) denote the unique


solution to the Neumann problem

(1 − ∆)u = f in Ω,
∂u
=0 on ∂Ω .
∂N
Show that K : L2 (Ω) → H 2 (Ω) is bounded, and K : L2 (Ω) → L2 (Ω) is
compact and symmetric. Construct an orthonormal basis {ek }∞ 2
k=1 in L (Ω)
which is orthogonal in H 1 (Ω) using the eigenfunctions of K.
Let u ∈ H 1 (Ω). Since u ∈ L2 (Ω), we may express u in terms of the basis
k
{e` }∞
P
`=1 . Suppose that uk = (u, e` )L2 (Ω) e` is the partial sum. Show that
`=1
uk → u in H 1 (Ω), and

kuk kH 1 (Ω) ≤ kukH 1 (Ω) . (5.75)

Suppose that u ∈ H 2 (Ω). In general uk does not converge to u since if so,


then
∂uk ∂u
0= → as k → ∞,
∂N ∂N
∂u
while does not have to vanish on ∂Ω. Now suppose that u ∈ H 2 (Ω)
∂N
∂u
satisfies = 0 on ∂Ω. Show that uk → u in H 2 (Ω), and there is a
∂N
constant C > 0 such that

kuk kH 2 (Ω) ≤ CkukH 2 (Ω) . (5.76)


Chapter 6

Linear Parabolic Equations

6.1 Spaces involving time

It is often important to identify functions u : Ω × [0, T ] → R with maps from


[0, T ] into a Banach space. For example, in the study of parabolic equations,
the space L2 (0, T ; H 1 (Ω)) plays an especially important role.
Suppose that X denotes an abstract Banach space with norm k · kX .

Definition 6.1. The space Lp (a, b; X) consists of functions u : (a, b) → X


so that u(t) ∈ X for almost all t ∈ (a, b). The Lp (a, b; X)-norm of u is given
as follows:
 hZ b i1
ku(·, t)kpX dt
p
if 1 ≤ p < ∞ ,



kukLp (a,b;X) ≡ a


 sup ku(·, t)kX if p = ∞ ,
t∈(a,b)

We say that Lp (a, b; X) is the space of Lp -integrable functions from (a, b)


into X.

Definition 6.2. The space C([a, b]; X) consists of functions u : [a, b] → X


continuously, so that u(t) ∈ X for all t ∈ (a, b). The C([a, b]; X)-norm of u
is given as follows:

kukC([a,b];X) = max ku(·, t)kX .


t∈[a,b]

For u ∈ C([a, b]; X), lim ku(·, t) − u(·, t0 )kX = 0 for all t0 ∈ [a, b]. In
t→t0
other words, C([a, b]; X) is the space of continuous functions from [a, b] into
X.

241
242 CHAPTER 6. Linear Parabolic Equations

Theorem 6.3. Lp (a, b; X) and C([a, b]; X) are Banach spaces for all p ∈
[1, ∞].

Proof. The proof is the same as the for the case that X = R. 

Lemma 6.4. Let X be a given Banach space with dual X 0 . Suppose that
u, g ∈ L1 (a, b; X). Then, the following three conditions are equivalent

(1) u is a.e. equal to the anti-derivative of g; that is, for some ξ ∈ X



sometimes denoted by u(a) ,
Z t
u(t) = ξ + g(s) ds for almost all t ∈ [a, b] , (6.1)
a

(2) For each test function ϕ ∈ D(a, b),


Z b Z b  dϕ 
u(t)ϕ 0 (t) dt = − g(t)ϕ(t) dt ϕ0 =
a a dt

(3) For each η ∈ X 0 ,


d
hu, ηiX = hg, ηiX ,
dt
in the sense of distribution.

If (1)-(3) are satisfied, u is equal to a continuous function from [a, b] into


X.

Proof. It is easy to see that (1) implies (2) and (3), while the proof of (2)
implies (1) is similar to the proof of Theorem 2.112 and is left as an exercise.
It suffices to show that (3) implies (2).
Suppose that (3) holds and ϕ ∈ D(a, b), then by definition
Z b Z b
hu(t), ηiX ϕ 0 (t) dt = − hg(t), ηiX ϕ(t) dt
a a

or
DZ b Z b E
u(t)ϕ 0 (t) dt + g(t)ϕ(t) dt, η =0 ∀η ∈ X0
a a X

which implies (2). 

Definition 6.5. The function g in (6.1) is called the weak time-derivative


of u, and is denoted by u 0 , ut or ∂t u.
6.1. SPACES INVOLVING TIME 243

The following corollary is the direct consequence of Lemma 6.4.

Corollary 6.6. If u 0 ∈ L1 (a, b; X), then u ∈ C([a, b]; X).

The notion of weak continuity plays a very important role in the analysis
of time-dependent PDE.

Definition 6.7. A function u ∈ L1 (a, b; X) is said to be weakly continuous


with values in X , denoted by u ∈ C([0, T ]; Xw ), if

hu, ϕiX ∈ C([0, T ]) ∀ϕ ∈ X0.

We use hu, ϕiX to denote the duality pairing between ϕ ∈ X 0 and u ∈ X.

Lemma 6.8. Let X and Y be two Banach spaces such that X ,→ Y (X


is continuously embedded in Y ) and X is dense in Y . If a function u ∈
L∞ (0, T ; X) and is weakly continuous with values in Y , then u is weakly
continuous with values in X.

Proof. Let ũ : R → X denote an extension of u to R satisfying ũ(t) =


u(t)χ[0,T ] (t), and define um = η 1 ∗ ũ where {η 1 }∞
m=1 denotes a sequence of
m m
standard mollifiers, and where the convolution operator is smoothing ũ in
time.
By definition, kum (t)kX ≤ kukL∞ (0,T ;X) for all t ∈ [0, T ], and since
u ∈ C([0, T ]; Yw ),

lim hum (t), ϕiX = hu(t), ϕiX ∀ t ∈ [0, T ] , ϕ ∈ Y 0 .


m→∞

Since hum (t), ϕiX ≤ kukL∞ (0,T ;X) kϕkX 0 for all m and t ∈ [0, T ], we find
that

∀ t ∈ [0, T ] , ϕ ∈ Y 0 .

hu(t), ϕiX = lim hum (t), ϕiX ≤ kukL∞ (0,T ;X) kϕkX 0
m→∞

This implies that the linear functional hu(t), ·iX is bounded on Y 0 . Since
Y 0 is dense in X 0 , this functional is also bounded on X 0 with

ku(t)kX = sup hu(t), ϕiX ≤ kukL∞ (0,T ;X) ∀ t ∈ [0, T ] .
kϕkX 0 =1
244 CHAPTER 6. Linear Parabolic Equations

Now suppose that t0 ∈ [0, T ] and ϕ ∈ X 0 . Then for any given  > 0, there
exists ψ ∈ Y 0 such that kϕ − ψkX 0 <  and we find that

hu(t) − u(t0 ), ϕiX ≤ hu(t) − u(t0 ), ψiX + hu(t) − u(t0 ), ϕ − ψiX

≤ hu(t) − u(t0 ), ψiY + 2kukL∞ (0,T ;X) .

Therefore, by the weak continuity of u with values in Y ,

lim sup |hu(t) − u(t0 ), ϕiX | ≤ 2kukL∞ (0,T ;X)


t→t0

which guarantees the weak continuity of u with values in X. 

Lemma 6.9. Suppose u ∈ L2 (0, T ; H), with u 0 ∈ L2 (0, T ; H 0 ) for some


Hilbert space H so that H ,→L2 (Ω) ⊆ H 0 (for example, H can be H 1 (Ω) or
H01 (Ω)). Then
d
(1) kuk2L2 (Ω) = 2hu0 , uiH in D0 ([0, T ]), where u0 is the weak time-derivative
dt
of u; and

(2) u ∈ C([0, T ]; L2 (Ω)), and


 
1 h i
max ku(t)kL2 (Ω) ≤ 1 + kukL2 (0,T ;H) + ku 0 kL2 (0,T ;H 0 ) . (6.2)
t∈[0,T ] T

Proof. Let ϕ ∈ C0∞ ([0, T ]), and uk ∈ C ∞ (Ω × [0, T ]) so that uk → u in


L2 (0, T ; H) and uk0 → u 0 in L2 (0, T ; H 0 ) as k → ∞. Then (1) follows from
the following equality:
Dd E Z TZ
0
kuk2L2 (Ω) , ϕ 2
|u(x, t)|2 ϕ 0 (t)dxdt


= − kukL2 (Ω) , ϕ = −
dt 0 Ω
Z TZ
= − lim |uk (x, t)|2 ϕ 0 (t)dxdt
k→∞ 0 Ω
Z TZ Z T
0
= lim 2uk (x, t)uk (x, t)ϕ(t)dxdt = 2hu 0 , uiH ϕdt .
k→∞ 0 Ω 0

To prove (2), we first show that u : [0, T ] → L2 (Ω) is continuous; that


is, we show that for each t0 ∈ [0, T ], lim ku(t) − u(t0 )k2L2 (Ω) = 0. However,
t→t0
lim ku(t)k2L2 (Ω) = ku(t0 )k2L2 (Ω) by (1) , and u(t) − u(t0 ), ϕ L2 (Ω) → 0 as

t→t0
t → t0 for all ϕ ∈ L2 (Ω) by Corollary 6.6 and Lemma 6.8 (with X = L2 (Ω)
6.1. SPACES INVOLVING TIME 245

and Y = H 0 ); hence

lim ku(t) − u(t0 )k2L2 (Ω)


t→t0
h i
= lim ku(t)k2L2 (Ω) + ku(t0 )k2L2 (Ω) − 2 u(t), u(t0 ) L2 (Ω) = 0 .

t→t0

Finally, by (1) we find that for almost all t, s ∈ [0, T ],


Z t
2 2
ku(t)kL2 (Ω) − ku(s)kL2 (Ω) = 2hu(s), u 0 (s)iH ds
s
≤ kuk2L2 (0,T ;H) + ku 0 k2L2 (0,T ;H 0 ) .

Integrate in s from 0 to T , the equality above implies


1 T
Z
ku(t)k2L2 (Ω) ≤ ku(s)k2L2 (Ω) ds + kuk2L2 (0,T ;H) + ku 0 k2L2 (0,T ;H 0 )
T 0
 
1 h i
≤ 1+ kuk2L2 (0,T ;H) + ku 0 k2L2 (0,T ;H 0 ) . 
T
Remark 6.10. The application of Corollary 6.6 and Lemma 6.8 for ob-
taining the continuity of u can be re-sentenced as “convergence in a weaker
sense (continuity in the weak sense) and the convergence of norms imply
convergence in a stronger sense (continuity in norm)”. This is a general-
ized version of the following well-known fact: if fn → f a.e. or weakly in
Lp (Ω), and kfn kLp (Ω) → kf kLp (Ω) , then fn → f in Lp (Ω).

Corollary 6.11. Let Ω ⊆ Rn be an open, bounded domain with smooth


boundary ∂Ω. Suppose that u ∈ L2 (0, T ; H m+1 (Ω)) and u 0 ∈ L2 (0, T ; H m−1 (Ω))
for some non-negative integer m. Then u ∈ C([0, T ]; H m (Ω)) and
h i
max ku(t)kH m (Ω) ≤ C kukL2 (0,T ;H m+1 (Ω)) + ku 0 kL2 (0,T ;H m−1 (Ω)) . (6.3)
t∈[0,T ]

6.1.1 Compactness theorems

By the Rellich theorem, H 1 (Ω)⊂⊂L2 (Ω) if Ω is bounded and smooth. How-


ever, it is not the case that L2 (0, T ; H 1 (Ω))⊂⊂L2 (0, T ; L2 (Ω)) since a se-
quence of functions uk (t) ∈ L2 (0, T ) clearly also belong to L2 (0, T ; H 1 (Ω))
but might not have convergent subsequence in L2 (0, T ; L2 (Ω)). In this sub-
section, we are concerned with pre-compact subsets of L2 (0, T ; H k (Ω)) or
L2 (0, T ; H01 (Ω)). .
246 CHAPTER 6. Linear Parabolic Equations

Lemma 6.12. Let X0 , X, X1 be Banach spaces such that

X0 ⊂⊂X ,→X1 . (6.4)

Then for each  > 0, there exists a constant C (and the spaces) such that

kxkX ≤ kxkX0 + C kxkX1 ∀ x ∈ X0 .

Proof. Suppose the contrary. Then for some fixed  > 0, there exists xk ∈ X0
such that

kxk kX ≥ kxk kX0 + kkxk kX1 .

Let yk = xk /kxk kX . Then the inequality above implies kkyk kX1 +kyk kX0 ≤
1. In particular, yk is bounded in X0 , and yk converges to 0 in X1 . By
X0 ⊂⊂X, there exists a subsequence ykj that converges to y in X. Since
yk converges to 0 in X1 , y must be 0 (this is the place where we use the
continuous embedding of X into X1 ), while the strong convergence of yk in
X implies that kykX = 1 which is a contradiction. 

Theorem 6.13. Let X0 , X, X1 be Banach spaces satisfying (6.4), and


X0 , X1 are reflexive. Let T > 0 be a fixed finite number, p0 , p1 ∈ (1, ∞),
and Y be the function space consisting of all u ∈ Lp0 (0, T ; X0 ) such that
u 0 ∈ Lp1 (0, T ; X1 ); that is,
n o
Y ≡ u ∈ L (0, T ; X0 ) u 0 ∈ Lp1 (0, T ; X1 )
p0

with norm kykY = kykLp0 (0,T ;X0 ) + kykLp1 (0,T ;X1 ) (which makes Y a Banach
space). Then

Y ⊂⊂Lp0 (0, T ; X) .

Proof. Let uk be a bounded sequence in Y . Since p0 , p1 ∈ (1, ∞) and X0 ,


X1 are reflexive, the spaces L2 (0, T ; X0 ) and L2 (0, T ; X1 ) are also reflexive;
hence Y is reflexive as well. By Banach-Alaoglu theorem, there exists ukj
that converges weakly to u ∈ Y ; that is,

ukj * u in Lp0 (0, T ; X0 ) and ukj t * ut in Lp1 (0, T ; X1 ) .


6.2. SECOND-ORDER PARABOLIC EQUATIONS 247

We show that this weak limit u is in fact the strong limit of a subsequence
of ukj in Lp0 (0, T ; X).
Let vj = ukj − u. The weak convergence of vjt in Lp1 (0, T ; X1 ) implies
that the function vj is a bounded sequence in C([0, T ]; X1 ) since by Lemma
6.4
Z t2 Z t2
kvj (t2 ) − vj (t1 )kX1 = vjt (s) ds ≤ kvjt (s)kX1 ds

t1 X1 t1
p1 −1
≤ |t2 − t1 | p1 kvjt kLp1 (0,T ;X1 ) .

The proof of the Arzela-Ascoli Theorem then also shows that there is a
further subsequence vj` (t) converges to 0 in C([0, T ]; X1 ). Hence vj` → 0 in
Lp0 (0, T ; X1 ).
Finally, by Lemma 6.12,

kvj` kLp0 (0,T ;X) ≤ kvj` kLp0 (0,T ;X0 ) + C kvj` kLp0 (0,T ;X1 ) . (6.5)

Therefore, the uniform boundedness of vj` in Lp0 (0, T ; X0 ) implies

lim kvj` kLp0 (0,T ;X) ≤ C ;


`→∞

hence vj` → 0 in Lp0 (0, T ; X0 ) or ukj` → u in Lp0 (0, T ; X0 ). 

6.2 Second-order parabolic equations

We consider the following initial-boundary value problem

ut + Lu = f in Ω × (0, T ) , (6.6a)
u = u0 on Ω × {t = 0} , (6.6b)

with either the Dirichlet boundary condition

u=0 on ∂Ω × (0, T ) (6.6c)

or the Neumann boundary condition

∂u
aij Ni = g on ∂Ω × (0, T ) , (6.6d)
∂xj
248 CHAPTER 6. Linear Parabolic Equations

where L is a time-dependent elliptic operator defined by

∂  ij ∂u  ∂u
(Lu)(x, t) = − a (x, t) (x, t) + bi (x, t) (x, t) (6.7)
∂xi ∂xj ∂xi
+ c(x, t)u(x, t)

and the matrix A = [aij ]. We note that since now problem is time dependent,
the coefficients a, b and c may depend on t as well.

Definition 6.14. We say that the partial differential operator + L is
∂t
uniformly parabolic if there exists a constant λ 0 > 0 such that aij (x, t)ξj ξi ≥
λ 0 |ξ|2 for all (x, t) ∈ Ω × (0, T ), ξ ∈ Rn .

6.2.1 Weak solutions

Assume that aij , bi , c ∈ L∞ (Ω × (0, T )), f ∈ L2 (0, T ; L2 (Ω)), and u0 ∈


L2 (Ω).

Definition 6.15 (Weak solutions with Dirichlet boundary conditions). A


function u ∈ L2 (0, T ; H01 (Ω)) with ut ∈ L2 (0, T ; H −1 (Ω)) is said to be a weak
solution of the parabolic initial-boundary value problem (6.6a,b,c) provided
that
 
ut (t), ϕ H 1 (Ω) + B u(t), ϕ = f (t), ϕ L2 (Ω)
0
(6.8)
∀ ϕ ∈ H01 (Ω), a.e. t ∈ (0, T ),
and

u(0) = u0 , (6.9)

where B u(t), ϕ is defined by

∂u ∂ϕ
Z h
aij (x, t)

B u(t), ϕ ≡ (x, t) (x)
Ω ∂xj ∂xi
(6.10)
i ∂u i
+ b (x, t) (x, t)ϕ(x) + c(x, t)u(x, t)ϕ(x) dx .
∂xi

Equation (6.8) is called the variational formulation of the parabolic problem


(6.6a,b,c).

Definition 6.16 (Weak solutions with Neumann boundary conditions). A


function u ∈ L2 (0, T ; H 1 (Ω)) with ut ∈ L2 (0, T ; H 1 (Ω)0 ) is said to be a weak
§6.2 Second-order parabolic equations 249

solution of the parabolic initial-boundary value problem (6.6a,b,d) provided


that u satsifies both (6.9) and

 

ut (t), ϕ H 1 (Ω) + B u(t), ϕ = f (t), ϕ L2 (Ω) + g(t), ϕ H 0.5 (∂Ω)
0
(6.11)
∀ ϕ ∈ H 1 (Ω), a.e. t ∈ (0, T ),

where B u(t), ϕ is defined by (6.10). Equation (6.11) is called the varia-
tional formulation of the parabolic problem (6.6a,b,d).

Remark 6.17. By Lemma 6.9, a weak solution u ∈ C([0, T ]; L2 (Ω)), so


condition (6.9) makes sense; in other words, the temporal trace at t = 0 is
well-defined.

We proceed to give the detailed construction of weak solutions for the


case of the Neumann boundary conditions (6.6d). The argument for the
case of the Dirichlet boundary condition (6.6c) is similar, and perhaps even
easier.

6.2.2 Existence of weak solutions - Galerkin schemes

Let {ek }∞ 2 1
k=1 be an orthonormal basis of L (Ω) which is orthogonal in H (Ω)
(given by Problem 5.9 in Chapter 5, for example). We define
k
X
uk (x, t) = dk` (t)e` (t)
`=1

such that the coefficients dk (t) are solutions of the following system:
  

ukt (t), ϕ L2 (Ω) + B uk (t), ϕ = f (t), ϕ L2 (Ω) + g(t), ϕ H 0.5 (∂Ω)
(6.12)
∀ ϕ ∈ span(e1 , · · · , ek ) .

The function uk (x, t) should be thought of as a finite-dimensional ap-


proximation of u(x, t); in fact, it is the L2 (Ω) projection onto the first k
orthonormal basis vectors {e1 , ... , el }.
The variational problem (6.12) is a coupled system of linear ordinary
differential equations for the vector dk (t) given by

dkm0 (t) + M`m (t)dk` (t) = (f (t), em )L2 (Ω) + g(t), em H 0.5 (∂Ω) ≡ Fm (t) ,

dk` (0) = (u0 , e` )L2 (Ω) ,


250 CHAPTER 6. Linear Parabolic Equations

where the time-dependent matrix M`m (t) is defined by


∂e` ∂em
Z h
M`m (t) = aij (x, t) (x) (x)
Ω ∂xj ∂xi
∂e` i
+ bi (x, t) (x)em (x) + c(x, t)e` (x)em (x) dx .
∂xi
This system of linear ordinary differential equations can be written as dk 0 (t)+
M̃(t)dk (t) = F (t); M̃(t) = MT (t) is a k × k matrix, and dk (t), F (t) are vec-
tors.
Since a, b, c ∈ L∞ ((0, T ); L∞ (Ω)), the existence of a unique Lipschitz
continuous functions dk is guaranteed by the fundamental theorem of ODE,
since the matrix M̃ ∈ L∞ (0, T ). To be more precise, by the contraction
mapping theorem, for each k ∈ N, there exists a 0 < Tk ≤ T such that the
solution dk (t) satisfies
Z t Z t
k k k
d (t) = d (0) − M̃(s)d (s) ds + F (s) ds ∀ t ∈ [0, Tk ] .
0 0

However, since the ODE is linear, Tk = T .


To obtain a weak solution to equation (6.6a,b,d), it is natural to consider
the limit of uk as k → ∞. To guarantee the existence of a weak limit, it is
necessary for us to obtain uniform estimates for the sequences uk and ukt ,
with bounds that are independent of k.

Uniform estimates for uk

Since (6.12) holds for all ϕ in the span of e1 , · · · , ek , we may set

ϕ = uk

in (6.12) since it is linear combination of {e` }k`=1 . By doing so, we find that
1d ∂uk ∂uk
Z Z
2 ij
kuk (t)kL2 (Ω) + a (x, t) (x, t) (x, t)dx= f (x, t)uk (x, t)dx
2 dt Ω ∂xj ∂xi Ω
∂u
Z h i
k
bi (x, t) (x, t)uk (x, t) + c(x, t)u2k (x, t) dx .


+ g(t), uk (t) H 0.5 (∂Ω) −
Ω ∂xi
By the uniform parabolicity, the equality above suggests that
1d
Z
2 2
kuk (t)kL2 (Ω) + λ 0 kDuk (t)kL2 (Ω) ≤ f (x, t)uk (x, t)dx
2 dt Ω
∂uk
Z h i
bi (x, t) (x, t)uk (x, t) + c(x, t)u2k (x, t) dx .


+ g(t), uk (t) H 0.5 (∂Ω) −
Ω ∂xi
§6.2 Second-order parabolic equations 251

Hölder’s and Young’s inequalities further imply that


1d
kuk (t)k2L2 (Ω) + λ 0 kDuk (t)k2L2 (Ω) ≤ kuk (t)kL2 (Ω) ×
2 dt h i
× kf (t)kL2 (Ω) + kb(t)kL∞ (Ω) kDuk (t)kL2 (Ω) + kc(t)kL∞ (Ω) kuk (t)kL2 (Ω)
λ0 h 1 1 λ0 i
≤ kDuk (t)k2L2 (Ω) + kb(t)k2L∞ (Ω) + kc(t)kL∞ (Ω) + + kuk (t)k2L2 (Ω)
2 λ0 2 4
1 C
+ kf (t)k2L2 (Ω) + kg(t)k2H −0.5 (∂Ω) ;
2 λ0
hence
d 2C
kuk (t)k2L2 (Ω) + λ 0 kDuk (t)k2L2 (Ω) ≤ kf (t)k2L2 (Ω) + kg(t)k2H −0.5 (∂Ω)
dt λ0
h 2 λ0 i
+ kb(t)k2L∞ (Ω) + 2kc(t)kL∞ (Ω) + 1 + kuk (t)k2L2 (Ω) .
λ0 2
| {z }
≡K(t)

By the generalized Gronwall inequality (see Problem 3.8 in Chapter 3),


h 2C i
kuk (t)k2L2 (Ω) ≤ kuk (0)k2L2 (Ω) + kf k2L2 (0,T ;L2 (Ω)) + kgk2L2 (0,T ;H −0.5 (∂Ω)) ×
λ0
Z t 
× exp K(s) ds
0
h i
≤ C ku0 k2L2 (Ω) + kf k2L2 (0,T ;L2 (Ω)) + kgk2L2 (0,T ;H −0.5 (∂Ω)) ,

for some constant C = C kbkL∞ (Ω×(0,T )) , kckL∞ (Ω×(0,T )) , λ 0 , T , where the
last inequality follows from the Bessel inequality stating that kuk (0)kL2 (Ω) ≤
ku0 kL2 (Ω) . As a consequence, dk exists on [0, T ], and
Z t
2
kuk (t)kL2 (Ω) + λ 0 kDuk (s)k2L2 (Ω) ds
i (6.13)
0
h
2 2 2
≤ C ku0 kL2 (Ω) + kf kL2 (0,T ;L2 (Ω)) + kgkL2 (0,T ;H −0.5 (∂Ω)) .

In other words, uk exists and is uniformly bounded in L2 (0, T ; H 1 (Ω)).

The estimates for ukt

To estimate kukt kH 1 (Ω)0 , we need to find the supremum of hukt , ϕiH 1 (Ω)
over functions ϕ ∈ H 1 (Ω) with kϕkH 1 (Ω) = 1. For each ϕ ∈ H 1 (Ω) with
kϕkH 1 (Ω) = 1, we can decompose ϕ so that ϕ = ϕ1 +ϕ2 , where ϕ1 belongs to
the span of e1 , · · · , ek and (ϕ1 , ϕ2 )H 1 (Ω) = 0. Then kϕ1 kH 1 (Ω) ≤ kϕkH 1 (Ω) =
1.
252 CHAPTER 6. Linear Parabolic Equations

Using the relation


  

ukt (t), ϕ1 L2 (Ω)
+ B uk (t), ϕ1 = f (t), ϕ1 L2 (Ω) + g(t), ϕ1 H 0.5 (∂Ω)

together with the fact that ukt belongs to the span of e1 , · · · , ek , we find
that

hukt (t), ϕiH 1 (Ω) = (ukt (t), ϕ)L2 (Ω) = (ukt (t), ϕ1 )L2 (Ω)


= (f (t), ϕ1 )L2 (Ω) + g(t), ϕ1 H 0.5 (∂Ω) − B uk (t), ϕ1 .

The expression for B uk (t), ϕ then shows that

B uk (t), ϕ1
h i
≤ ka(t)kL∞ (Ω) + kb(t)kL∞ (Ω) + kc(t)kL∞ (Ω) kuk (t)kH 1 (Ω) kϕ1 kH 1 (Ω) .

As a consequence,

kukt kH 1 (Ω)0 = sup hukt (t), ϕiH 1 (Ω) ≤ kf (t)kL2 (Ω) + kg(t)kH −0.5 (∂Ω)
kϕkH 1 (Ω) =1
h i
+ ka(t)kL∞ (Ω) + kb(t)kL∞ (Ω) + kc(t)kL∞ (Ω) kuk (t)kH 1 (Ω) ;

hence
Z T h i
kukt (t)k2H 1 (Ω)0 dt ≤ 4 kf k2L2 (0,T ;L2 (Ω)) + kgk2L2 (0,T ;H −0.5 (∂Ω))
0
Z Th i
+8 ka(t)k2L∞ (Ω) + kb(t)k2L∞ (Ω) + kc(t)k2L∞ (Ω) kuk (t)k2H 1 (Ω) dt .
0

By (6.13), the inequality above implies that


Z T h i
kukt (t)k2H 1 (Ω)0 dt ≤ C |u0 k2L2 (Ω) + kf k2L2 (0,T ;L2 (Ω)) + kgk2L2 (0,T ;H −0.5 (∂Ω))
0

for some constant C = C kakL∞ (Ω×(0,T )) , kbkL∞ (Ω×(0,T )) , kckL∞ (Ω×(0,T )) , T, λ 0 );


thus ukt is uniformly bounded in L2 (0, T ; H 1 (Ω)0 ).
Since L2 (0, T ; H 1 (Ω)) and L2 (0, T ; H 1 (Ω)0 ) are both Hilbert spaces, by
the Banach-Alaoglu Theorem A.10 there exists a subsequence kj of k such
that ukj and ukj t converge weakly in L2 (0, T ; H 1 (Ω)) and L2 (0, T ; H 1 (Ω)0 ),
respectively. Let u be the weak limit of ukj ; then the weak limit of ukj t must
be ut .
§6.2 Second-order parabolic equations 253

N
d˜` (t)e` (x) for some N ≤ kj , d˜` ∈ C 1 ([0, T ]). Since
P
Let v(x, t) =
`=1
v(t) ∈ span(e1 , · · · , ekj ), (6.12) implies that

 

ukj t (t), v(t) H 1 (Ω) + B ukj (t), ϕ = f (t), v(t) L2 (Ω) + g(t), v(t) H 0.5 (∂Ω) .

Integrating the equality above in time over [0, T ], the definition of weak
convergence in L2 (0, T ; H 1 (Ω)) and L2 (0, T ; H 1 (Ω)0 ) shows that
Z Th

i
ut (t), v(t) H 1 (Ω) + B u(t), v(t) dt
0
Z Th i


= f (t), v(t) L2 (Ω) + g(t), v(t) H 0.5 (∂Ω) dt .
0
N
d˜` (t)e` (x), is dense in L2 (0, T ; H 1 (Ω)), the equality
P
Since v, of the form
`=1
above implies that
Z Th Z Th

i 
ut (t), v(t) H 1 (Ω) + B u(t), v(t) dt = f (t), v(t) L2 (Ω)
0
i
0 (6.14)
+ g(t), v(t) H 0.5 (∂Ω) dt ∀ v ∈ L (0, T ; H 1 (Ω)) .
2

Finally, letting v(x, t) = χ[a,b] (t)ϕ(x) for some ϕ ∈ H 1 (Ω) in (6.14), we


find that for 0 ≤ a < b ≤ T ,
Z bh

i
ut (t), ϕ H 1 (Ω) + B u(t), ϕ dt
a
Z bh i
f (t), ϕ L2 (Ω) + g(t), ϕ H 0.5 (∂Ω) dt ∀ ϕ ∈ H 1 (Ω),


=
a
(6.15)
and the variational formulation (6.11) follows from the Lebesgue differenti-
ation theorem.

Remark 6.18. From the discussion above, we know that (6.11) and (6.14)
are equivalent, so (6.14) can be thought as an alternative variational formu-
lation for the parabolic initial-boundary value problem (6.6a,b,d). Similarly,
an alternative variational formulation for equation (6.6a,b,c) is given by
Z Th

i
ut (t), v(t) H 1 (Ω) + B u(t), v(t) dt
0
Z T
∀ v ∈ L2 (0, T ; H01 (Ω)) .

= f (t), v(t) L2 (Ω) dt
0
(6.16)
254 CHAPTER 6. Linear Parabolic Equations

In order to prove that u is indeed a weak solution, it remains to prove


that u(0) = u0 . We note that while by construction, uk (0) → u0 in L2 (Ω),
it is not necessary that u(0) = u0 .
Let ζ ∈ C 1 [0, T ] be such that ζ(0) = 1 and ζ(T ) = 0. Multiplying


(6.12) by ζ(t) and integrating over the time interval (0, T ), we find that for
all ϕ ∈ span(e1 , · · · , ekj ),
Z T h  i
ukj t (t), ζ(t)ϕ L2 (Ω)
+ B ukj (t), ζ(t)ϕ dt
0
Z T h 
i
= f (t), ζ(t)ϕ L2 (Ω) + g(t), ζ(t)ϕ H 0.5 (∂Ω) dt ,
0
(6.17)
and the use of v(x, t) = ζ(t)ϕ(x) (which belongs to L2 (0, T ; H 1 (Ω))) in (6.14)
implies that
Z Th

i
ut (t), ζ(t)ϕ L2 (Ω) + B u(t), ζ(t)ϕ dt
0
Z T h 
i
= f (t), ζ(t)ϕ L2 (Ω) + g(t), ζ(t)ϕ H 0.5 (∂Ω) dt ,
0
(6.18)
for all ϕ ∈ H 1 (Ω). Integrating-by-parts in time on the first term of the
left-hand side of (6.17), we find that
Z Th i
B ukj (t), ζ(t)ϕ − ukj (t), ζ 0 (t)ϕ L2 (Ω) dt (6.19)
  
ukj (0), ϕ H 1 (Ω) +
0
Z Th i


= f (t), ϕ L2 (Ω) + g(t), ζ(t)ϕ H 0.5 (∂Ω) dt ∀ ϕ ∈ span(e1 , · · · , ekj ) ,
0

and similarly from (6.18),


Z Th i
B u(t), ζ(t)ϕ − u(t), ζ 0 (t)ϕ L2 (Ω) dt
  
u(0), ϕ L2 (Ω) + (6.20)
0
Z Th i
f (t), ϕ L2 (Ω) + g(t), ζ(t)ϕ H 0.5 (∂Ω) dt ∀ ϕ ∈ H 1 (Ω).


=
0

Passing (6.19) to the limit, since ukj (0) → u0 in L2 (Ω), we obtain that
T  Z
B u(t), ζ(t) − hu(t), ζ 0 (t)ϕiH 1 (Ω) dt
 
(u0 , ϕ)L2 (Ω) + (6.21)
0
Z Th i
f (t), ϕ L2 (Ω) + g(t), ζ(t)ϕ H 0.5 (∂Ω) dt ∀ ϕ ∈ H 1 (Ω).


=
0
§6.2 Second-order parabolic equations 255


The comparison of (6.20) and (6.21) suggests that (u0 , ϕ)L2 (Ω) = u(0), ϕ L2 (Ω)
for all ϕ ∈ H 1 (Ω); hence u(0) = u0 a.e. This proves the existence of a weak
solution to equation (6.6a,b,d).

6.2.3 Uniqueness of the weak solution

Assume that u1 and u2 are two weak solutions. Let w = u1 − u2 . Then by


(6.14) w satisfies
Z T h
i
∀ v ∈ L2 (0, T ; H 1 (Ω)) ,

wt (t), v(t) H 1 (Ω) + B w(t), v(t) dt = 0
0
w=0 at t = 0 .

Since w ∈ L2 (0, T ; H 1 (Ω)), we may use it in the equality above, and similar
to the estimates that we obtained for uk , we find that
t=T Z T
kw(t)k2L2 (Ω) + λ0 kDw(t)k2L2 (Ω) dt

t=0 0
h kbk iZ T
L∞ (0,T ;L∞ (Ω))
≤ + 2kckL∞ (0,T ;L2 (Ω)) kw(t)k2L2 (Ω) dt .
2λ 0 0

The Gronwall inequality then implies that kw(t)k2L2 (Ω) = 0 since w(0) = 0,
hence the uniqueness of the weak solution to equation (6.6a,b,d).
We have established the following

Theorem 6.19. There exists a unique weak solution u to the parabolic


initial-boundary value problem (6.6a,b,c) (resp. (6.6a,b,d)) satisfying

max ku(t)kL2 (Ω) + kukL2 (0,T ;H 1 (Ω)) + kut kL2 (0,T ;H −1 (Ω))
t∈[0,T ]
h i
≤ C ku0 kL2 (Ω) + kf kL2 (0,T ;L2 (Ω))
 
max ku(t)kL2 (Ω) + kukL2 (0,T ;H 1 (Ω)) + kut kL2 (0,T ;H 1 (Ω)0 )
t∈[0,T ]
resp.
 
h i
≤ C ku0 kL2 (Ω) + kf kL2 (0,T ;L2 (Ω)) + kgkL2 (0,T ;H −0.5 (∂Ω))
(6.22)
for some constant C depending on λ 0 , T , kakL∞ (0,T ;L∞ (Ω)) , kbkL∞ (0,T ;L∞ (Ω))
and kckL∞ (0,T ;L∞ (Ω)) .

Remark 6.20. The condition f ∈ L2 (0, T ; L2 (Ω)) can be further weakened


to f ∈ L2 (0, T ; H 1 (Ω)0 ), and the right-hand side of the weak formulation
256 CHAPTER 6. Linear Parabolic Equations

(6.11) is then replaced by hf (t), ϕiH01 (Ω) or hf (t), ϕiH 1 (Ω) . The estimates are

almost identical, and the right-hand side of (6.22) will then be C ku0 kL2 (Ω) +
  
kf kL2 (0,T ;H −1 (Ω)) or C ku0 kL2 (Ω) + kf kL2 (0,T ;H 1 (Ω)0 ) .

Remark 6.21. Unlike the elliptic case, no matter how large b and c are,
the solution to the parabolic initial-boundary value problem (6.6a,b,c) or
(6.6a,b,d) always exists and is unique. For the problem with the Neumann
boundary conditions, the initial condition is used to rule out the non-zero
constant functions in the kernel of the differential operator ∂t + L.

6.2.4 Regularity

The regularity result for the parabolic initial-boundary value problem relies
on elliptic regularity, and estimates for the time-derivatives of the solution.
For example, suppose that we could prove that ut (t) ∈ L2 (Ω); then according
to the elliptic regularity theorem for the Neumann boundary condition, we
conclude that u(t) ∈ H 2 (Ω).
Improving the regularity of ut requires more regularity of the initial
data and boundary data. To proceed, we assume that u0 ∈ H 1 (Ω), g ∈
L2 (0, T ; H 0.5 (∂Ω)) with gt ∈ L2 (0, T ; H −0.5 (∂Ω)), and the matrix A = aij
 

is symmetric; that is, aij = aji . Setting ϕ = ukt in (6.12), we find that

∂uk ∂ukt
Z
kukt (t)k2L2 (Ω) + aij (x, t)
(x, t) (x, t)dx
Ω ∂xj ∂xi
Z Z
= f (x, t)ukt (x, t)dx + g(x, t)ukt (x, t)dS
Ω ∂Ω
∂uk
Z h i
− bi (x, t) (x, t)ukt (x, t) + c(x, t)uk (x, t)ukt (x, t) dx .
Ω ∂xi

Since we assume that a is symmetric,

∂uk ∂ukt 1d ∂uk ∂uk


Z Z
aij (x, t) (x, t) (x, t)dx = aij (x, t) (x, t) (x, t)dx
Ω ∂xj ∂xi 2 dt Ω ∂xj ∂xi
1 ∂uk ∂uk
Z
− aij
t (x, t) (x, t) (x, t) dx . (6.23)
2 Ω ∂xj ∂xi

3 2 1 2
Hölder’s inequality together with Young’s inequality, ab ≤ a + b , then
2 6
§6.2 Second-order parabolic equations 257

leads to
1d ∂uk ∂uk
Z
kukt (t)k2L2 (Ω) + aij (x, t)
(x, t) (x, t)dx
2 dt
Ω ∂xj ∂xi
1 3 h
≤ kat (t)kL∞ (Ω) kDuk (t)k2L2 (Ω) + kb(t)k2L∞ (Ω) kDuk (t)k2L2 (Ω)
2 2 i 1
+ kc(t)k2L∞ (Ω) kuk (t)k2L2 (Ω) + kf (t)k2L2 (Ω) + kukt (t)k2L2 (Ω)
Z 2
+ g(x, t)ukt (x, t)dS .
∂Ω
(6.24)
For the last term on the right-hand side, we note that integrating by parts
in time implies that
Z tZ
g(x, s)ukt (x, s) dSds
0 ∂Ω
Z s=t Z t

= g(x, s)uk (x, s)dS − gt (s), uk (s) H 0.5 (∂Ω) dt

∂Ω s=0 0
≤ Ckg(t)kH −0.5 (∂Ω) kuk (t)kH 1 (Ω) + Ckg(0)kH −0.5 (∂Ω) kuk (0)kH 1 (Ω)
Z t
+C kgt (s)kH −0.5 (∂Ω) kuk (s)kH 1 (Ω) ds
0
h 1
≤ C ku0 k2H 1 (Ω) + kg(0)k2H −0.5 (∂Ω) + kg(t)k2H −0.5 (∂Ω) (6.25)
4λ 0
i λ
0
+ kgt k2L2 (0,T ;H −0.5 (∂Ω)) + kf kL2 (0,T ;L2 (Ω)) + kuk (t)k2H 1 (Ω) .
2

Moreover, by the interpolation inequality


h i
max kg(t)kL2 (∂Ω) ≤ C kgkL2 (0,T ;H 0.5 (∂Ω)) + kgt kL2 (0,T ;H −0.5 (∂Ω)) , (6.26)
t∈[0,T ]

(6.25) implies that


Z tZ h
g(x, s)ukt (x, s) dSds ≤ C ku0 k2H 1 (Ω) + kf k2L2 (0,T ;L2 (Ω))
0 ∂Ω
i λ
0
+ kgk2L2 (0,T ;H 0.5 (∂Ω)) + kgt k2L2 (0,T ;H −0.5 (∂Ω)) + kuk (t)k2H 1 (Ω) .
2
(6.27)
As a consequence, integrating (6.24) over the time interval (0, t), by the
parabolicity and (6.13) we find that
t
λ0
Z
kukt (s)k2L2 (Ω) ds + kDuk (t)k2L2 (Ω)
0 2
h i
≤ C ku0 k2H 1 (Ω) + kf k2L2 (Ω) + kgk2L2 (0,T ;H 0.5 (∂Ω)) + kgt k2L2 (0,T ;H −0.5 (∂Ω)
258 CHAPTER 6. Linear Parabolic Equations

for some constant C = C kakL∞ (Ω×(0,T )) , kbkL∞ (Ω×(0,T )) , kckL∞ (Ω×(0,T )) , T, λ 0 ).


Therefore, the weak limit ut belongs to L2 (0, T ; L2 (Ω)) and satisfies
h
kut k2L2 (0,T ;L2 (Ω)) ≤ C ku0 k2H 1 (Ω) + kf k2L2 (0,T ;L2 (Ω)) + kgk2L2 (0,T ;H 0.5 (∂Ω))
i
+ kgt k2L2 (0,T ;H −0.5 (∂Ω)) .

Next, we rewrite the weak formulation (6.11) as

B(u, ϕ) = (f − ut , ϕ)L2 (Ω) + (g, ϕ)L2 (∂Ω) ∀ ϕ ∈ H 1 (Ω), t ∈ (0, T ) a.e. .

For almost every instant of time, we thus have a weak solution u to the ellip-
∂u
tic problem Lu = f − ut with the Neumann boundary condition aij Ni =
∂xj
g. By elliptic regularity,
h i
ku(t)kH 2 (Ω) ≤ C(t) kf (t) − ut (t)kL2 (Ω) + kgkH 0.5 (∂Ω) + ku(t)kL2 (Ω)

for some constant C(t) = C ka(t)kC 1 (Ω) , kb(t)kC 1 (Ω) , kc(t)kC 1 (Ω) , λ 0 . Thus
for a, b, c ∈ C 1 (Ω × [0, T ]),
h
kukL2 (0,T ;H 2 (Ω)) ≤ C ku0 kH 1 (Ω) + kf kL2 (0,T ;L2 (Ω)) + kgkL2 (0,T ;H 0.5 (∂Ω))
i
+ kgt kL2 (0,T ;H −0.5 (∂Ω)) .

We conclude with the following

Theorem 6.22. Assume that ∂t +L is uniformly parabolic, ∂Ω is of class C 2 ,


aij , bi , c ∈ C 1 (Ω × [0, T ]), and aij is symmetric. Then for any u0 ∈ H 1 (Ω),
 

f ∈ L2 (0, T ; L2 (Ω)), g ∈ L2 (0, T ; H 0.5 (∂Ω)) with gt ∈ L2 (0, T ; H −0.5 (∂Ω)),


the unique weak solution u to (6.6a,b,d) in fact belongs to L2 (0, T ; H 2 (Ω)) ∩
C([0, T ]; H 1 (Ω)) with ut ∈ L2 (0, T ; L2 (Ω)) and satisfies

max ku(t)kL2 (Ω) + kut kL2 (0,T ;L2 (Ω)) + kukL2 (0,T ;H 2 (Ω))
t∈[0,T ]
h
≤ C ku0 kH 1 (Ω) + kf kL2 (0,T ;L2 (Ω)) + kgkL2 (0,T ;H 0.5 (∂Ω)) (6.28)
i
+ kgt kL2 (0,T ;H −0.5 (∂Ω)) .

Remark 6.23. For the Dirichlet problem (6.6a,b,c), the same argument
leading to Theorem 6.22 can be applied to obtain ut ∈ L2 (0, T ; L2 (Ω)) with
§6.2 Second-order parabolic equations 259

the additional assumption u0 = 0 on ∂Ω. The reason for requiring that u0


vanishes on ∂Ω is for the validity of the following inequality

kDuk (0)kL2 (Ω) ≤ Cku0 kH 1 (Ω) (6.29)

for some constant C independent of k, and u0 = 0 on ∂Ω is called the


first-order compatibility condition for the parabolic initial-boundary problem
(6.6a,b,c). We note that if u0 does not vanish on ∂Ω, (6.29) is no longer
true (see Problem 6.6 for the counterexample). Thus similar to Theorem
6.22 we have the following

Theorem 6.24. Assume that ∂t +L is uniformly parabolic, ∂Ω is of class C 2 ,


aij , bi , c ∈ C 1 (Ω × [0, T ]), and aij is symmetric. Then for any u0 ∈ H01 (Ω),
 

f ∈ L2 (0, T ; L2 (Ω)), the unique weak solution u to (6.6a,b,c) in fact belongs


to L2 (0, T ; H 2 (Ω)) ∩ C([0, T ]; H01 (Ω)) with ut ∈ L2 (0, T ; L2 (Ω)) and satisfies

max ku(t)kL2 (Ω) + kut kL2 (0,T ;L2 (Ω)) + kukL2 (0,T ;H 2 (Ω))
t∈[0,T ]
  (6.30)
≤ C ku0 kH 1 (Ω) + kf kL2 (0,T ;L2 (Ω)) .

6.2.5 Higher regularity and compatibility conditions

To address the issues of obtaining higher regularity, we first focus the dis-
cussion on one of the simplest parabolic equations

ut − ∆u = f in Ω × (0, T ) , (6.31a)
∂u
=g on ∂Ω × (0, T ) , (6.31b)
∂N
u = u0 on Ω × {t = 0} . (6.31c)

Suppose in addition that f ∈ L2 (0, T ; H 1 (Ω)) with ft ∈ L2 (0, T ; H 1 (Ω)0 ),


g ∈ L2 (0, T ; H 1.5 (∂Ω)), and u0 ∈ H 2 (Ω). A solution of the time-differentiated
problem

wt − ∆w = ft in Ω × (0, T ) , (6.32a)
∂w
= gt on ∂Ω × (0, T ) , (6.32b)
∂N
w = f (0) + ∆u0 ≡ u1 on Ω × {t = 0} , (6.32c)
260 CHAPTER 6. Linear Parabolic Equations

can be related to the solution of (6.31) via the fundamental theorem of


calculus; that is,
Z t
u(x, t) = u0 (x) + w(x, s) ds ∀x ∈ Ω. (6.33)
0

(We have assumed sufficient regularity on u0 such that w|t=0 ∈ L2 (Ω).)


To validate (6.33), we consider the weak form of (6.32):




wt (t), ϕ H 1 (Ω) + Dw(t), Dϕ L2 (Ω) = ft (t), ϕ H 1 (Ω) + gt (t), ϕ H 0.5 (∂Ω)
∀ ϕ ∈ H 1 (Ω), a.e. t ∈ (0, T ).

Integrating the equation above in time, we find that for all ϕ ∈ H 1 (Ω),

Z t 
w(t), ϕ H 1 (Ω) + Dw(s) ds, Dϕ 2
L (Ω)

0


= w(0), ϕ H 1 (Ω) + f (t) − f (0), ϕ H 1 (Ω) + g(t) − g(0), ϕ H 0.5 (∂Ω)
 
= f (t), ϕ L2 (Ω) + (∆u0 , ϕ)L2 (Ω) + g(t) − g(0), ϕ L2 (∂Ω)
 
= f (t), ϕ L2 (Ω) + g(t), ϕ L2 (∂Ω) − (Du0 , Dϕ)L2 (Ω)
∂u0
Z h i
+ − g(0) ϕ dS ;
∂Ω ∂N

hence for all ϕ ∈ H 1 (Ω),


 Z t Z t
    
u0 + w(s)ds t , ϕ + D u0 + w(s) ds , Dϕ
0 L2 (Ω) 0 L2 (Ω)
∂u0
Z h i
 
= f (t), ϕ L2 (Ω) + g(t), ϕ L2 (∂Ω) + − g(0) ϕ dS .
∂Ω ∂N
Z t
∂u0
This shows that u0 + w(s) ds is a weak solution to (6.31) if = g(0),
0 ∂N
and by the uniqueness of the weak solution, we indeed have (6.33). The
condition
∂u0
= g(0) on ∂Ω (6.34)
∂N
is called the first-order compatibility condition for (6.31).
Now since ft ∈ L2 (0, T ; H 1 (Ω)0 ) and u0 ∈ H 2 (Ω), the weak solution
w = ut to (6.32) belongs to L2 (0, T ; H 1 (Ω)) with wt ∈ L2 (0, T ; H(Ω)0 )
satisfying

max kut (t)kL2 (Ω) + kut kL2 (0,T ;H 1 (Ω)) + kutt kL2 (0,T ;H(Ω)0 )
t∈[0,T ]
h i
≤ C ku0 kH 2 (Ω) + kf (0)kL2 (Ω) + kft kL2 (0,T ;H(Ω)0 ) + kgt kL2 (0,T ;H −0.5 (∂Ω)) .
§6.2 Second-order parabolic equations 261

Moreover, −∆u = f − ut in the weak sense, so elliptic regularity implies


that
h i
ku(t)kH 3 (Ω) ≤ C kf (t)kH 1 (Ω) + kut (t)kH 1 (Ω) + kgkH 1.5 (∂Ω)

and hence
h i
max kut (t)kL2 (Ω) + ku(t)kH 2 (Ω) + kukL2 (0,T ;H 3 (Ω)) + kut kL2 (0,T ;H 1 (Ω))
t∈[0,T ]
h
+ kutt kL2 (0,T ;H(Ω)0 ) ≤ C ku0 kH 2 (Ω) + kf kL2 (0,T ;H 1 (Ω)) + kft kL2 (0,T ;H(Ω)0 )
i
+ kgkL2 (0,T ;H 1.5 (∂Ω)) + kgt kL2 (0,T ;H −0.5 (∂Ω))

for some constant C = C(T ). Here we use (6.2) to estimate kf (0)kL2 (Ω) .
On the other hand, if u ∈ L2 (0, T ; H 3 (Ω)) with ut ∈ L2 (0, T ; H 1 (Ω)),
∂u
then u ∈ C([0, T ]; H 2 (Ω)) thus Du ∈ C([0, T ]; H 1 (Ω)) which implies ∈
∂N
∂u
C([0, T ]; H 0.5 (∂Ω)). Since by construction = g on ∂Ω, we have
∂N
∂u0 ∂u(t)
= lim = g(0) on ∂Ω .
∂N t→0 ∂N
Therefore, (6.34) is also a necessary condition for (6.31) to have a solution
u ∈ L2 (0, T ; H 3 (Ω)) with ut ∈ L2 (0, T ; H 1 (Ω)).
Suppose further that u0 ∈ H 3 (Ω) and that f ∈ L2 (0, T ; H 2 (Ω)), ft ∈
L2 (0, T ; L2 (Ω)), g ∈ L2 (0, T ; H 2.5 (∂Ω)), gt ∈ L2 (0, T ; H 0.5 (∂Ω)) and gtt ∈
L2 (0, T ; H −0.5 (∂Ω)). Then the solution w to the time-differentiated problem
(6.32) belongs to L2 (0, T ; H 2 (Ω)) ∩ L∞ (0, T ; H 1 (Ω)) by Theorem 6.22. No
additional compatibility condition is required to guarantee the additional
gain of regularity of w. However, in order to make a connection between w
and u via (6.33), the first-order compatibility condition is still needed.
Suppose even further that u0 ∈ H 4 (Ω) and that f ∈ L2 (0, T ; H 3 (Ω)),
ft ∈ L2 (0, T ; H 1 (Ω)), and ftt ∈ L2 (0, T ; H 1 (Ω)0 ), as well as g ∈ L2 (0, T ; H 3.5 (∂Ω)),
gt ∈ L2 (0, T ; H 1.5 (∂Ω)) and gtt ∈ L2 (0, T ; H −0.5 (∂Ω)). In this case, w, ft ,
gt and u1 play the role of u, f , g and u0 in (6.31); hence to understand the
regularity of the solution, it is necessary to time-differentiate (6.32) once
more. We find that the solution u to (6.31) belongs to L2 (0, T ; H 5 (Ω)) as
long as one additional condition on the initial data is satisfied:

∂w(0) ∂ f (0) + ∆u0
= = gt (0) on ∂Ω . (6.35)
∂N ∂N
262 CHAPTER 6. Linear Parabolic Equations

The condition (6.35) is the first-order compatibility condition for the time-
differentiated problem (6.32), and is called the second-order compatibility
condition for (6.31).
In the following, we summarize what we have concluded in a table.

k The requirement for u∈L2 (0, T ;H k (Ω)) in addition to u0 ∈H k−1 (Ω)

1 f ∈L2 (0, T ;H 1 (Ω)0 ), g ∈L2 (0, T ;H −0.5(∂Ω)).

2 f ∈L2 (0, T ; L2 (Ω)), g ∈L2 (0, T ;H 0.5(∂Ω)), gt ∈L2 (0, T ;H −0.5(∂Ω)).

f ∈L2 (0, T ;H 1 (Ω)), ft ∈L2 (0, T ;H 1 (Ω)0 ), g ∈L2 (0, T ;H 1.5(∂Ω)),


3
gt ∈L2 (0, T ;H −0.5(∂Ω)), 1st-order compatibility condition.

f ∈L2 (0, T ;H 2 (Ω)), ft ∈L2 (0, T ; L2 (Ω)), g ∈L2 (0, T ;H 2.5(∂Ω)),


4 gt ∈L2 (0, T ;H 0.5(∂Ω)), gtt ∈L2 (0, T ;H −0.5(∂Ω)),
1st-order compatibility condition.

f ∈L2 (0, T ;H 3 (Ω)), ft ∈L2 (0, T ;H 1 (Ω)), ftt ∈L2 (0, T ;H 1 (Ω)0 ),
5 g ∈L2 (0, T ;H 3.5(∂Ω)), gt ∈L2 (0, T ;H 1.5(∂Ω)), gtt ∈L2 (0, T ;H −0.5(∂Ω)),
1st-order and 2nd-order compatibility conditions.

The discuss above motivates the following

Definition 6.25 (Compatible time-differentiated problems). The k-th com-


patible time-differentiated version of the parabolic initial-boundary value prob-
lem (6.6a,b,c) (resp. (6.6a,b,d)) consists of the following:
∂ku
1. a parabolic equation for wk ≡ obtained from time-differentiating
∂tk
(6.6a) k times ;

2. an initial condition defined by

∂ k−1
uk ≡ wk (0) = − (f − Lu) ; (6.36)
∂tk−1 t=0

3. a Dirichlet (resp. Neumann) boundary condition defined by


k−1
∂kg
ij ∂wk
X k ∂ k−` aij ∂w` 
 
wk = 0 resp. a Ni = k − Ni on ∂Ω.
∂xj ∂t ` ∂tk−` ∂xj
`=0
(6.37)
§6.2 Second-order parabolic equations 263

We note that when considering equation (6.6a,b,d), the corresponding


boundary condition for the k-th compatible time-differentiated problem is
∂u
obtained from time-differentiating the boundary condition aij Ni = g k
∂xj
times.

Definition 6.26 (Compatibility conditions). For k ∈ N ∪ {0}, the (k + 1)th-


order compatibility condition for the parabolic initial-boundary value problem
(6.6a,b,c) (resp. (6.6a,b,d)) is given by τ uk = 0 on ∂Ω
k−1 
∂uk ∂kg X k ∂ k−` aij ∂u`
  
resp. aij (0) Ni = k (0) − Ni (0) on ∂Ω ,
∂xj ∂t ` ∂tk−` ∂xj
`=0

where u` , ` = 1, · · · , k, is the initial condition of the `th compatible time-


differentiated problem for equation (6.6a,b,c) (resp. (6.6a,b,d)) defined in
(6.36).

Finally, for the purpose of defining the regularity of the solution and the
k (T ; Ω) and V k (T ; Ω)
forcing functions, we introduce the function spaces VD N
that consist of space-time dependent functions so that one time derivative
scales like two space derivatives. To be more concrete, for k ≥ 1,

VDk (T ; Ω)
n h k + 1 io
≡ u ∈ L (0, T ; H (Ω)) ∂tj u ∈ L2 (0, T ; H k−2j (Ω)) if 0 ≤ j ≤
2 k
,

2

[ k+1
2
]
k∂tj ukL2 (0,T ;H k−2j (Ω)) , and
X
equipped with norm kukV k (T ;Ω) =
D
j=0

VNk (T ; Ω)
n hki
≡ u ∈ L2 (0, T ; H k (Ω)) ∂tj u ∈ L2 (0, T ; H k−2j (Ω)) if 0 ≤ j ≤ ,

2
h k i [ k ]+1 o
k−2 ∂t 2 u ∈ L2 (0, T ; H 1 (Ω)0 ) ,
2

equipped with norm


k
[2]
[ k ]+1
h k i
k∂tj ukL2 (0,T ;H k−2j (Ω)) + k − 2
X
kukV k (T ;Ω) = k∂t 2 ukL2 (0,T ;H 1 (Ω)0 ) .
N 2
j=0

k (T ; Ω) and V k (T ; Ω) are the same when k is even.


Note that VD N
264 CHAPTER 6. Linear Parabolic Equations

k+0.5
For the Neumann problem, we also need the space V∂Ω (T ) for defining
the regularity of the boundary data g: for k ∈ N ∪ {0, −1},

k+0.5
V∂Ω (T )
n hk + 1i
≡ g ∈L2 (0, T ; H k+0.5 (∂Ω)) ∂tj g ∈L2 (0, T ; H k+0.5−2j (∂Ω)) if 0 ≤ j ≤

o 2
k/2+1 2 −0.5
∂t g ∈ L (0, T ; H (∂Ω)) .

equipped with norm

[ k+1
2
]
k∂tj gkL2 (0,T ;H k+0.5−2j (∂Ω))
X
kgkV k+0.5 (T ) =
∂Ω
j=0
[ k+3
h k + 1 i ]
+ k+1−2 k∂t 2
gkL2 (0,T ;H −0.5 (∂Ω)) .
2

Having defined the function spaces, we can state the main

Theorem 6.27. Assume that ∂Ω is of class C m+2 , ∂t + L is uniformly


parabolic, aij , bi , c ∈ C m+1 (Ω × [0, T ]), and aij is symmetric. Then for
 

any u0 ∈ H m+1 (Ω) ∩ H01 (Ω) (resp. H m+1 (Ω)), f ∈ VDm (T ; Ω) (resp. f ∈
m+0.5
VNm (T ; Ω), g ∈ V∂Ω (T )), the unique weak solution u to the parabolic
initial-boundary value problem (6.6a,b,c) (resp. (6.6a,b,d)) in fact belongs to
VDm+2 (T ; Ω) (resp. VNm+2 (T ; Ω)) and satisfies
h i
kukV m+2 (T ;Ω) ≤C ku0 kH m+1 (Ω) + kf kVDm (T ;Ω)
D
 h i (6.38)
resp. kukV m+2 (T ;Ω) ≤C ku0 kH m+1 (Ω) + kf kVNm (T ;Ω) + kgkV m+0.5 (T ) ,
N ∂Ω

hmi 
provided that the compatibility conditions are valid up to + 1 -th (resp.
hm + 1i 2
-th) order.
2

6.3 Parabolic equations with special forcing func-


tions

We will explore a kind of parabolic initial-boundary value problem which


involves special kind of forcing functions and is useful in many real world
applications (and one of the applications will be seen in Chapter 7). The
§6.3 Parabolic equations with special forcing functions 265

equation we are going to study is

∂  ij ∂u  ∂  ij ∂h 
ut − a =f+ a in Ω × (0, T ) , (6.39a)
∂xi ∂xj ∂xi ∂xj
∂u ∂h
aij Ni = g − aij Ni on ∂Ω × (0, T ) , (6.39b)
∂xj ∂xj
u = u0 in Ω × {t = 0} , (6.39c)

where aij , ∈ L∞ (Ω × (0, T )), u0 ∈ L2 (Ω), f ∈ L2 (0, T ; L2 (Ω)) and g ∈


−0.5
V∂Ω (T ) as before, and in addition we assume that aij ∈ L∞ (Ω×(0, T )) and
h ∈ L2 (0, T ; H 1 (Ω)). Here we set bi = c = 0 in L to simplify the discussion.
At the first sight, (6.39) does not look very different from the parabolic
problem (6.6a,b,d) if we treat (aij h,j ),i as the forcing function and (6.39b)
is merely a non-homogeneous Neumann boundary condition. This is exactly
true when h possesses better regularity, say for example h ∈ L2 (0, T ; H 2 (Ω));
∂h
however, with h ∈ L2 (0, T ; H 1 (Ω)), the boundary data g−aij Ni does not
∂xj
make sense, so requires a little special treatment. Moreover, this problem is
of special interest especially when study the regularity theory for hyperbolic
equations.
Testing (6.39a) against a test function ϕ ∈ H 1 (Ω) results in the following

Definition 6.28. A function u ∈ L2 (0, T ; H 1 (Ω)) with ut ∈ L2 (0, T ; H 1 (Ω)0 )


is said to be a weak solution to (6.39) if
Z

  
ut (t), ϕ H 1 (Ω) + B u(t), ϕ = f (t), ϕ L2 (Ω) − B h(t), ϕ + g(t)ϕ dS
∂Ω
∀ ϕ ∈ H 1 (Ω), a.e. t ∈ (0, T ) , (6.40)

and u(0) = u0 , where B(u, ϕ) is defined by (6.10) with bi = c = 0, and


B(h, ϕ) is defined similarly with aij replacing aij . (6.40) is called the vari-
ational formulation of the parabolic equation (6.39).

For f, g, h with given regularity, (6.40) makes perfect sense, and the
existence of the unique weak solution to (6.39) follows exactly the same
fashion as before. In fact, since


B(h, ϕ) ≤ CkDhkL2 (Ω) kDϕkL2 (Ω) ,
266 CHAPTER 6. Linear Parabolic Equations

by Young’s inequality, the finite dimensional approximation uk satisfies


1d λ0
kuk k2L2 (Ω) + kDuk k2L2 (Ω)
2 dt 2
1 Ch i 1 λ0 
≤ kf k2L2 (Ω) + kDhk2L2 (Ω) + kgk2H −0.5 (∂Ω) + + kuk k2L2 (Ω)
2 λ0 2 4
which by Gronwall’s inequality gives uniform bounds for uk in L2 (0, T ; H 1 (Ω))
and ukt in L2 (0, T ; H 1 (Ω)0 ); thus the weak limit provides the unique weak
solution to (6.39).
Now suppose in addition that u0 ∈ H 1 (Ω), aij , aij ∈ C 1 (Ω × [0, T ]),
[aij ] is symmetric, g ∈ L2 (0, T ; H 0.5 (∂Ω)) and h ∈ L2 (0, T ; H 2 (Ω)). Then
Theorem 6.22 implies that u ∈ VN 2 (T ; Ω) and satisfies

Z Th i
2
max ku(t)kH 1 (Ω) + ku(t)k2H 2 (Ω) + kut (t)k2L2 (Ω) dt
t∈[0,T ] 0
h i
≤ C ku0 kH 1 (Ω) + kf k2L2 (0,T ;L2 (Ω)) + kgk2L2 (0,T ;H 0.5 (∂Ω)) + khk2L2 (0,T ;H 2 (Ω)) .
2

Suppose further that u0 ∈ H 2 (Ω), aij , aij ∈ C 2 (Ω × [0, T ]), f ∈ V 1 (T ; Ω),


1.5 (T ), and h ∈ L2 (0, T ; H 3 (Ω)) with h ∈ L2 (0, T ; H 1 (Ω)). For better
g ∈ V∂Ω t

regularity of u, we need to study the time differentiated problem


∂  ij ∂w  ∂  ij ∂ht ∂h ∂u 
wt − a = ft + a + aij
t + aij
t in Ω×(0, T ),
∂xi ∂xj ∂xi ∂xj ∂xj ∂xj
∂w
h ∂ht ∂h ∂u i
aij Ni = gt − aij + aij
t + aij
t Ni on ∂Ω×(0, T ),
∂xj ∂xj ∂xj ∂xj
∂  ij ∂u0  ∂  ij ∂h(0) 
w = f (0) − a − a on Ω×{t = 0}.
∂xi ∂xj ∂xi ∂xj

By Corollary 6.11, h ∈ C([0, T ]; H 2 (Ω)) thus h(0) ∈ H 2 (Ω) which implies


that w(0) ∈ L2 (Ω). Therefore, the time-differentiated problem is in the form
of (6.39), and possesses a unique weak solution satisfying
Z Th i
2
max kw(t)kL2 (Ω)+ kw(t)k2H 1 (Ω)+kwt (t)k2H 1 (Ω)0 dt
t∈[0,T ] 0
h
≤C kf (0)k2L2 (Ω)+kh(0)k2H 2 (Ω)+ku0 k2H 2 (Ω)+kft k2L2 (0,T ;H 1 (Ω)0 )+kuk2L2 (0,T ;H 1 (Ω))
i
+khk2L2 (0,T ;H 1 (Ω))+kht k2L2 (0,T ;H 1 (Ω))+kgt k2L2 (0,T ;H −0.5 (∂Ω))
h
≤C ku0 k2H 2 (Ω)+kf k2L2 (0,T ;H 1 (Ω))+kft k2L2 (0,T ;H 1 (Ω)0 )+khk2L2 (0,T ;H 3 (Ω))
(6.41)
i
+kht k2L2 (0,T ;H 1 (Ω))+kgt k2L2 (0,T ;H −0.5 (∂Ω)) .
6.4. EXERCISES 267

As long as the first-order compatibility condition


∂u0 ∂h(0)
aij (0) Ni = g(0) − aij (0) Ni on ∂Ω × (0, T )
∂xj ∂xj

holds, (6.33) is satisfied and (6.41) implies that


h i
kuk2V 3 (T ;Ω) ≤ C ku0 k2H 2 (Ω) + kf k2V 1 (T ;Ω) + kgk2V 1.5 (T ) + khk2V 3 (T ;Ω) .
N N ∂Ω N

Inductively, by Theorem 6.27 we establish the following

Theorem 6.29. Assume that ∂Ω is of class C m+2 , ∂t + L is uniformly


parabolic, aij , aij , bi , c ∈ C m+1 (Ω × [0, T ]), and aij is symmetric. Then for
 

m (T ; Ω), g ∈ V m+0.5 (T ), h ∈ V m+2 (T ; Ω), the


any u0 ∈ H m+1 (Ω), f ∈ VN ∂Ω N
unique weak solution u to the parabolic initial-boundary value problem
∂  ij ∂h 
ut + Lu = f + a in Ω × (0, T ) ,
∂xi ∂xj
∂u ∂h
aij Ni = g − aij Ni on ∂Ω × (0, T ) ,
∂xj ∂xj
u = u0 in Ω × {t = 0} ,

m+2 m+2
in fact belongs to VN (T ; Ω) (resp. VN (T ; Ω)) and satisfies
h
kukV m+2 (T ;Ω) ≤ C ku0 kH m+1 (Ω) + kf kVNm (T ;Ω) + kgkV m+0.5 (T )
N ∂Ω
i (6.42)
+ khkV m+2 (T ;Ω) ,
N

hm + 1i
provided that the compatibility conditions are valid up to -th order.
2

6.4 Exercises

Problem 6.1. Assume that uj * u in L2 (0, T ; H01 (Ω)) and that ∂t uk * v


in L2 (0, T ; H −1 (Ω)). Prove that v = ∂t u. (Hint. Let φ ∈ C01 (0, T ), and
Z T Z T
1
w ∈ H0 (Ω). Then h∂t uk , φwidt = − huk , φwidt.)
0 0

Problem 6.2. Let H be a Hilbert space, and u, ut ∈ L2 (0, T ; H). Show that
u ∈ C([0, T ]; H) and
 1 h i
max ku(t)k2H ≤ 1 + kuk2L2 (0,T ;H) + kut k2L2 (0,T ;H) . (6.43)
t∈[0,T ] T
268 CHAPTER 6. Linear Parabolic Equations

Problem 6.3. Let g ∈ C([0, T ]; L2 (Ω)) and u be such that ut ∈ L2 (0, T ; H 2 (Ω))
and
−κ∆ut − ∆u = g on Ω × [0, T ].

Then, independently of κ > 0,

k∆ukL∞ (0,T ;L2 (Ω)) ≤ kgkL∞ (0,T ;L2 (Ω)) + k∆u(0)kL2 (Ω) .

Problem 6.4. Let Ω = (0, 1)2 and let

1
(Ω) := u ∈ H 1 (Ω) u = 0 on {y = 0} ∩ {x = 0} ∩ {x = 1}

HB

Prove the existence of a weak solution u ∈ L2 (0, T ; HB


1 (Ω) ∩ H 1 (ω)) to the

parabolic problem

ut − ∆u = f in Ω × (0, T ] ,
u=0 on {x = 0} ∩ {x = 1} ∩ {y = 0} × (0, T ] ,
uy = uxx on (ω := {y = 1}) × (0, T ] ,
u = u0 on Ω × {t = 0} ,

where f ∈ L2 (0, T ; L2 (Ω) and u0 ∈ L2 (Ω).


Discuss the requirements on the data for regularity of solutions.

Problem 6.5. Let Ω ⊂ Rd for d > 1, and consider the time-dependent


Stokes problem for the velocity vector u = (u1 , ..., ud ) and the scalar pressure
function p, given by

ut − ∆u + Dp = f in Ω,
divu = 0 in Ω, (6.44)
u=0 on ∂Ω

where f ∈ L2 (0, T ; L2 (Ω; Rd )). Let

1
(Ω) = u ∈ H01 (Ω; Rd ) divu = 0 .

H0,div

Prove that there exists a unique weak solution u ∈ L2 (0, T ; H0,div


1 (Ω)) to the
Stokes problem.
Discuss the requirements on the data for regularity of solutions.
§6.4 Exercise 269
r
2
Problem 6.6. Let e` = sin `x, and ẽ` = `−1 e` . From Chapter 3 we
π
know that {e` }∞
`=1 is an orthonormal basis of L2 (0, π). Let the inner product
on H01 (0, π) be defined by (u, v)H01 (0,π) = (u0 , v 0 )L2 (0,π) . By Corollary 5.67
{ẽ` }∞ 1
`=1 is an orthonormal basis of H0 (0, π). Show that

(u, e` )L2 (0,π) = `−1 (u 0 , ẽ`0 )L2 (0,π) ∀ u ∈ H 1 (0, π);

k k
(u, ẽ` )H01 (0,π) ẽ` converges to u in H 1 (0, π)
P P
thus uk = (u, e` )L2 (0,π) e` =
`=1 `=1
for all u ∈ H01 (0, π).
Let u = 1 be a constant function. Then
k
4 X 4 sin(2` + 1)x
u2k+1 =
π π 2` + 1
`=0

Show that there is no C > 0 independent of k such that

ku0k kL2 (0,π) ≤ CkukH 1 (0,π) = Cπ .


Chapter 7

Linear Hyperbolic Equations

We consider the initial-boundary value problem

utt + Lu = f in Ω × (0, T ) , (7.1a)


u = u0 on Ω × {t = 0} , (7.1b)
ut = u1 on Ω × {t = 0} , (7.1c)

with either the Dirichlet boundary condition

u=0 on ∂Ω × (0, T ) (7.1d)

or the Neumann boundary condition


∂u
aij Ni = g on ∂Ω × (0, T ) , (7.1e)
∂xj
where, as for the parabolic case, L is a time-dependent elliptic operator
defined by
∂  ij ∂u  ∂u
(Lu)(x, t) = − a (x, t) (x, t) + bi (x, t) (x, t) (7.2)
∂xi ∂xj ∂xi
+ c(x, t)u(x, t)

satisfying the uniform hyperbolicity condition that for some constant λ 0 > 0,

aij (x, t)ξj ξi ≥ λ 0 |ξ|2 for all (x, t) ∈ Ω × (0, T ), ξ ∈ Rn .

7.1 Weak solutions

We assume that the matrix A = [aij ] is symmetric; that is, aij = aji , and
aij , aij i 2 2 2
t , b , c ∈ C(Ω × [0, T ]), f ∈ L (0, T ; L (Ω)), g ∈ L (0, T ; H
−0.5 (∂Ω))

270
7.1. WEAK SOLUTIONS 271

with gt ∈ L2 (0, T ; H −0.5 (∂Ω)), u0 ∈ H 1 (Ω) and u1 ∈ L2 (Ω). We remark that


if aij , bi , c are time-independent, then only aij , bi , c ∈ L∞ (Ω) are required.

Definition 7.1 (Weak solutions with Dirichlet boundary conditions). A


function u ∈ L∞ (0, T ; H01 (Ω)) with time-derivatives ut ∈ L∞ (0, T ; L2 (Ω))
and utt ∈ L2 (0, T ; H −1 (Ω)) is said to be a weak solution of the hyperbolic
initial-boundary value problem (7.1a,b,c,d) provided that

 
utt (t), ϕ H 1 (Ω) + B u(t), ϕ = f (t), ϕ L2 (Ω)
0
(7.3)
∀ ϕ ∈ H01 (Ω), a.e. t ∈ (0, T ) ,

and

u = u0 , ut = u1 on Ω × {t = 0} , (7.4)

where B u(t), ϕ is defined by

∂u ∂ϕ
Z h
aij (x, t)

B u(t), ϕ ≡ (x, t) (x)
Ω ∂xj ∂xi
(7.5)
i ∂u i
+ b (x, t) (x, t)ϕ(x) + c(x, t)u(x, t)ϕ(x) dx .
∂xi
Equation (7.3) is called the variational formulation of the hyperbolic problem
(7.1a,b,c,d).

Definition 7.2 (Weak solutions with Neumann boundary conditions). A


function u ∈ L∞ (0, T ; H 1 (Ω)) with time-derivatives ut ∈ L∞ (0, T ; L2 (Ω))
and utt ∈ L2 (0, T ; H 1 (Ω)0 ) is said to be a weak solution of the hyperbolic
initial-boundary value problem (7.1a,b,c,e) provided that

 

utt (t), ϕ H 1 (Ω) + B u(t), ϕ = f (t), ϕ L2 (Ω) + g(t), ϕ H 0.5 (∂Ω)
0
(7.6)
∀ ϕ ∈ H01 (Ω), a.e. t ∈ (0, T ) ,

and (7.4) holds. Here again B u(t), ϕ is defined by (7.5). Equation (7.6)
is called the variational formulation of the hyperbolic problem (7.1a,b,c,e).

Remark 7.3. In order to establish the existence of the weak solution to the
hyperbolic equation, we already have to assume the symmetry of A, while
the same condition is used to prove the regularity theory of the parabolic
problem.
272 CHAPTER 7. Linear Hyperbolic Equations

Remark 7.4. Let u be a weak solution to the hyperbolic problem (7.1a,b,c,d).


By definition u ∈ C([0, T ]; L2 (Ω)) and ut ∈ C([0, T ]; H −1 (Ω)); thus we can
expect that

lim ku(t) − u0 kL2 (Ω) = lim kut (t) − u1 kH −1 (Ω) = 0. (7.7)


t→0 t→0

Moreover, by Lemma 6.8, u ∈ C([0, T ]; H01 (Ω)w ) and ut ∈ C([0, T ]; L2 (Ω)w );


that is,

lim hu(t) − u(t0 ), ϕiH01 (Ω) = 0 ∀ ϕ ∈ H −1 (Ω) , t0 ∈ [0, T ] ,


t→t0
∀ ϕ ∈ L2 (Ω) , t0 ∈ [0, T ] ,

lim ut (t) − ut (t0 ), ϕ L2 (Ω) = 0
t→t0

so in addition to (7.7), we also have

∀ ϕ ∈ H −1 (Ω),


lim u(t) − u0 , ϕ H 1 (Ω) = 0
t→0
 0
lim ut (t) − u1 , ϕ L2 (Ω) = 0 ∀ ϕ ∈ L2 (Ω).
t→0

Similar arguments apply to the hyperbolic problem (7.1a,b,c,e). Therefore,


(7.4) in the definition of weak solutions is understood as

lim ku(t) − u0 kL2 (Ω) = lim kut (t) − u1 kX 0 = 0, (7.8)


t→0 t→0

and
lim hu(t) − u0 , ϕiX = 0 ∀ ϕ ∈ X 0,
t→t0
(7.9)
∀ ϕ ∈ L2 (Ω),

lim ut (t) − u1 , ϕ L2 (Ω) = 0
t→t0

where X = H01 (Ω) for Definition 7.1 or X = H 1 (Ω) for Definition 7.2.

Remark 7.5. As for the parabolic case, the argument leading (6.14) to
(6.11) implies that the variational formulation (7.3) is equivalent to
Z Th

i
utt (t), ϕ(t) H 1 (Ω)+ B u(t), ϕ(t) dt
0
0
(7.10)
Z T
2 1

= f (t), ϕ(t) L2 (Ω)
dt ∀ ϕ ∈ L (0, T ; H (Ω)),
0

and the variational formulation (7.6) is equivalent to


Z Th

i
utt (t), ϕ(t) H 1 (Ω)+ B u(t), ϕ(t) dt (7.11)
0
Z Th i
f (t), ϕ(t) L2 (Ω)+ g(t), ϕ(t) L2 (∂Ω) dt ∀ ϕ ∈ L2 (0, T ; H 1 (Ω)).
 
=
0
7.1. WEAK SOLUTIONS 273

In the following, we focus only on the discussion of the hyperbolic equa-


tion with the Neumann boundary condition, while similar argument can be
made for the hyperbolic equation with the Dirichlet boundary condition.
We start with the uniqueness of the weak solution.

7.1.1 Uniqueness of the weak solution

Assume that u1 and u2 are two weak solutions. Let w = u1 − u2 . Then by


(7.11) w satisfies
Z Th
i
∀ v ∈ L2 (0, T ; H 1 (Ω)), (7.12a)


wtt (t), v(t) H 1 (Ω)+ B w(t), v(t) dt = 0
0
w = wt = 0 at t = 0. (7.12b)

Since w ∈ L∞ (0, T ; H 1 (Ω)), the function v defined by


 Z s
 w(τ )dτ if 0 ≤ t ≤ s ,
v(t) = t
0 if s < t ≤ T .

belongs to L2 (0, T ; H 1 (Ω)); thus v can be used as a test function in (7.12a).


Integrating by parts in time,

t=s Z sh  i
wt (t), v (t) H 1 (Ω) − wt (t), vt (t) L2 (Ω) − B w(t), v (t) dt = 0 .

t=0 0
(7.13)

Therefore, by Lemma 6.9 and (7.12b), v(s) = 0, vt = −w in [0, s],


t=s Z s 1 s d
Z
kw(t)k2L2 (Ω) dt.


wt (t), v(t) H 1 (Ω) − wt (t), vt (t) L2 (Ω) dt =

t=0 0 2 0 dt

By the symmetry of A = [aij ],

ij ∂vt ∂v 1d ij ∂v ∂v 1 ∂v ∂v
Z Z Z
a dx = a dx − aij
t dx .
Ω ∂xj ∂xi 2 dt Ω ∂xj ∂xi 2 Ω ∂xj ∂xi

Moreover, integrating by parts in t, by v(s) = 0 we obtain that


sZ sZ
i ∂vt ∂2v
Z Z
b vdxdt = bi v dxdt
0 Ω ∂xi 0 Ω ∂xi ∂t
Z sZ
∂v ∂v
Z
= bi (0)v(0) (0) dx − (bi v)t dx .
Ω ∂xi 0 Ω ∂xi
274 CHAPTER 7. Linear Hyperbolic Equations

As a consequence, (7.13) implies that


s
1 dh ∂v ∂v
Z Z i
2
kw(t)kL2 (Ω) − aij (t) (t) (t) dx dt
0 2 dt Ω ∂xj ∂xi
Z sh Z
1 ∂v ∂v ∂v
Z Z i
=− aij dx + (bi
v)t dx + cwvdx dt
0 2 Ω t ∂xj ∂xi Ω ∂xi Ω
∂v(0)
Z
+ bi (0) v(0) dx .
Ω ∂xi

Again by w(0) = 0 and v(s) = 0, as well as the uniform hyperbolicity


assumption,
Z sh
kw(s)k2L2 (Ω) + λ 0 kDv(0)k2L2 (Ω) ≤ kat (t)kL∞ (Ω) kDv(t)k2L2 (Ω)
0

+ kbt (t)kL∞ (Ω) kv(t)kL2 (Ω) + kb(t)kL∞ (Ω) kw(t)kL2 (Ω) kDv(t)kL2 (Ω)
i
+ kc(t)kL∞ (Ω) kw(t)kL2 (Ω) kv(t)kL2 (Ω) dt
λ0 1
+ kDv(0)k2L2 (Ω) + kb(0)k2L∞ (Ω) kv(0)k2L2 (Ω)
2 2λ 0
Z sh
≤ kat (t)kL∞ (Ω) + kbt (t)kL∞ (Ω) + kb(t)kL∞ (Ω) + kc(t)kL∞ (Ω)
0
s ih i λ0
kb(0)k2L∞ (Ω) kv(t)k2H 1 (Ω) + kw(t)k2L2 (Ω) dt +
+ kDv(0)k2L2 (Ω)
2λ 0 2
Z sh
≤ (1 + s) kat (t)kL∞ (Ω) + kbt (t)kL∞ (Ω) + kb(t)kL∞ (Ω) + kc(t)kL∞ (Ω)
0
kb(0)k2L∞ (Ω) ih i λ0
+ kDv(t)k2L2 (Ω) + kw(t)k2L2 (Ω) dt + kDv(0)k2L2 (Ω) .
2λ 0 2
Z t
Let W (t) = w(τ )dτ . Then the inequality above implies that
0

2kw(s)k2L2 (Ω) + λ 0 kDW (s)k2L2 (Ω)


Z sh i
≤ C(1+ s) kDW (t)− DW (s)k2L2 (Ω) + kw(t)k2L2 (Ω) dt
Z0 s h i
≤ C(1+ s) kDW (t)k2L2 (Ω) + kw(t)k2L2 (Ω) dt+ C(s + s2 )kDW (s)k2L2 (Ω)
0

or

2kw(s)k2L2 (Ω) + λ 0 − C(s + s2 ) kDW (s)k2L2 (Ω)



Z sh i
≤ C(1 + s) kDW (t)k2L2 (Ω) + kw(t)k2L2 (Ω) dt .
0
7.2. THE METHOD OF PARABOLIC REGULARIZATION - THE SIMPLE CASE275

For T1 > 0 small enough so that λ 0 − C(T1 + T12 ) ≥ λ 0 /2, the Gronwall
inequality implies that for all t ∈ [0, T1 ],
h i
kw(t)k2L2 (Ω) + kDW (t)k2L2 (Ω) ≤ C kw(0)k2L2 (Ω) + kDW (0)k2L2 (Ω) = 0 ;

hence w(t) = 0 or u1 (t) = u2 (t) in [0, T1 ]. In other words, the solution to


(7.1a,b,c,e) is unique in the time interval [0, T1 ]. The same argument applies
to the interval [T1 , 2T1 ], [2T1 , 3T1 ], etc., so the solution is unique in the time
interval [0, T ]. Therefore, we establish the following

Theorem 7.6. There exists at most one weak solution to the hyperbolic
initial-boundary value problem (7.1a,b,c,d) or (7.1a,b,c,e).

7.2 The method of parabolic regularization - the


simple case

Given the uniqueness of the weak solution, we only focus on the existence
and the regularity theory in the rest of the discussion. To illustrate the idea,
we start with the wave equation; that is, aij = δij and bi = c = 0, and will
return to the general case later.

7.2.1 The existence of the weak solution

We first study the existence of a weak solution solution to the hyperbolic


initial-boundary value problem

utt − ∆u = f in Ω × (0, T ) , (7.14a)


∂u
=g on ∂Ω × (0, T ) , (7.14b)
∂N
u = u0 , ut = u1 on Ω × {t = 0} . (7.14c)

We would like to prove the following

Theorem 7.7. There exists one and only one weak solution to the wave
equation (7.14).

It suffices to show the existence of a weak solution to (7.14). To prove


the existence, instead of considering (7.14) directly, we consider the following
276 CHAPTER 7. Linear Hyperbolic Equations

parabolic κ-problem

uκtt − κ∆uκt − ∆uκ = f in Ω × (0, T ) , (7.15a)


∂uκ ∂uκ
κ t + =g on ∂Ω × (0, T ) , (7.15b)
∂N ∂N
uκ = u0 , uκt = u1 on Ω × {t = 0} , (7.15c)

where κ > 0 is artificial viscosity introduced into the system, and look for the
(weak) limit of uκ as κ → 0. The superscript κ indicates that the solution
uκ depends on the parameter κ. We remark that by letting w = uκt , (7.15a)
is a parabolic equation of w (if ∆uκ is treated as a forcing function), and
(7.15) is called the zeroth-order parabolic regularization of (7.14).
One of the strategy of solving (7.15) for all fixed κ > 0 is through a
Z t
fixed-point argument: given w̃ ∈ L2 (0, T ; H 1 (Ω)), let ũ = u0 + w̃(s) ds,
0
and solve

wt − κ∆w = f + ∆ũ in Ω × (0, T ) , (7.16a)


∂w ∂ ũ
κ =g− on ∂Ω × (0, T ) , (7.16b)
∂N ∂N
w = u1 on Ω × {t = 0} . (7.16c)

By the discussion in Section 6.3 the weak solution to (7.16) exists; thus we
establish a map Φ : L2 (0, T ; H 1 (Ω)) → L2 (0, T ; H 1 (Ω)) by Φ(w̃) = w, and
a fixed-point of Φ is a weak solution to (7.15).
The existence of a fixed-point relies on some type of fixed-point theorem.
Here we will employee the famous contraction mapping theorem.

Lemma 7.8 (Contraction mapping theorem). Let X be a complete metric


space, and Ψ : X → X be a contraction mapping on X; that is, for some
constant 0 < C < 1,


dX Ψ(x), Ψ(y) ≤ CdX (x, y) ∀ x, y ∈ X ,

where dX denotes the metric on X. Then Φ has a unique fixed-point in X.

We show that Φ is a contraction mapping on L2 (0, T ; H 1 (Ω)) provided


that T > 0 is small enough. Now given w̃1 and w̃2 in L2 (0, T ; H 1 (Ω)), let
§7.2 The method of parabolic regularization - the simple case 277
Z t
ũi = u0 + w̃i (s) ds, δ ũ = ũ1 − ũ2 , wi = Φ(w̃i ), and δ w̃ = w̃1 − w̃2 . Then
0
δw = w1 − w2 satisfies δw(0) = 0, and


(δw)t , ϕ H 1 (Ω) + κ(Dδw, Dϕ)L2 (Ω) = Dδ ũ, Dϕ L2 (Ω)
(7.17)
∀ ϕ ∈ H 1 (Ω), a.e. t ∈ (0, T ) .

Since δw ∈ L2 (0, T ; H 1 (Ω)), δw(t) ∈ H 1 (Ω) for almost all t ∈ (0, T ); thus
can be used as a test function in (7.17). By doing so, we find that

1d
kδwk2L2 (Ω) + κkDδwk2L2 (Ω) = (Dδ ũ, Dδw)L2 (Ω)
2 dt
1 κ
≤ kDδ ũk2L2 (Ω) + kDδwk2L2 (Ω) .
2κ 2

Therefore, by δw(0) = 0,
t t
1
Z Z
kδw(t)k2L2 (Ω) + κ kDδw(s)k2L2 (Ω) ds ≤ kDδ ũ(s)k2L2 (Ω) ds
0 κ 0
Z t
Now, since δ ũ = δ w̃(s) ds,
0
Z t √ hZ t i1
2
kDδ ũkL2 (Ω) ≤ kDδ w̃(s)kL2 (Ω) ds ≤ t kDδ w̃(s)k2L2 (Ω) ds .
0 0

As a consequence,
Z t
kδw(t)k2L2 (Ω)
+κ kDδw(s)k2L2 (Ω) ds
0
1 t
Z s
t2 t
Z Z
0 2 0
≤ s kDδ w̃(s )kL2 (Ω) ds ds ≤ kDδ w̃(s)k2L2 (Ω) ds .
κ 0 0 2κ 0

Moreover, integrating in time, the energy estimate above implies that


t t
t3
Z Z
kδw(s)k2L2 (Ω) ds ≤ kDδ w̃(s)k2L2 (Ω) ds ;
0 6κ 0

thus
t h t3 t
t2 i
Z Z
kδw(s)k2H 1 (Ω) ≤ + 2 kDδ w̃(s)k2L2 (Ω) ds .
0 6κ 2κ 0

Note that the inequality above is the same as


r
t3 t2
kΦ(w̃1 ) − Φ(w̃2 )kL2 (0,T ;H 1 (Ω)) ≤ + 2 kw̃1 − w̃2 kL2 (0,T ;H 1 (Ω)) ; (7.18)
6κ 2κ
278 CHAPTER 7. Linear Hyperbolic Equations

hence the contraction mapping theorem implies that there is a Tκ > 0 such
that the unique weak solution to (7.15) exists in the time interval [0, Tκ ].
Having established the existence of the weak solution uκ to (7.15), the
next step is to passing κ to the limit which requires a κ-independent up-
per bounded for uκ ∈ L∞ (0, Tκ ; H 1 (Ω)), uκt ∈ L∞ (0, Tκ ; L2 (Ω)) and uκtt ∈
L2 (0, Tκ ; H 1 (Ω)0 ). Moreover, from the process of constructing the weak so-
lution, the time of existence Tκ appears to approach 0 as κ → 0 (so that
the contraction constant is strictly smaller than 1); thus in addition to a
uniform upper bound of uκ , we also need to prove that the time of existence
can be made independent of κ, then pass κ → 0 to obtain the weak solution
to (7.14).

κ-independent estimates
Z t
κ
The unique fixed point w satisfies u (t) = u0 + w(s)ds or uκt = w. There-
0
fore, we may rewrite the variational formulation of (7.16) as


κ
utt (t), ϕ H 1 (Ω) + κ Duκt (t), Dϕ L2 (Ω) + Duκ (t), Dϕ L2 (Ω)
 

= f (t), ϕ L2 (Ω) + g, ϕ H 0.5 (∂Ω) ∀ ϕ ∈ H 1 (Ω) , a.e. t ∈ (0, T ).




(7.19)
Since uκt = w ∈ L2 (0, Tκ ; H 1 (Ω)), it can be used as a test function, and we
obtain that

1 dh κ 2 i
kut kL2 (Ω) + kDuκ k2L2 (Ω) + κkDuκt k2L2 (Ω)
2 dt
1 1
≤ kf k2L2 (Ω) + kuκt k2L2 (Ω) + hg, uκt iH 0.5 (∂Ω) .
2 2

Similar to (6.27),

Z t s=t Z t
hg, uκt iH 0.5 (∂Ω) ds κ
hgt , uκ iH 0.5 (∂Ω) ds


= g(s), u (s) H 0.5 (∂Ω) −

0 s=0 0
h i
≤ C kg(0)k2H −0.5 (∂Ω) + ku0 k2H 1 (Ω) + C kg(t)k2H −0.5 (∂Ω) (7.20)
Z t
+ kuκ (t)k2H 1 (Ω) + Ckgt k2L2 (0,T ;H −0.5 (∂Ω)) + C kuκ (s)k2H 1 (Ω) ds.
0
§7.2 The method of parabolic regularization - the simple case 279

Therefore, integrating the inequality above in time over the time interval
(0, t) and choosing  > 0 small enough, we find that
Z t
kuκt (t)k2L2 (Ω) + kDu κ
(t)k2L2 (Ω) + κ kDuκt (s)k2L2 (Ω) ds
0
h
≤C kuκt (0)k2L2 (Ω) + kDu κ
kf k2L2 (0,T ;L2 (Ω)) + kgk2L2 (0,T ;H −0.5 (∂Ω))
(0)k2L2 (Ω) +
i Z th i
2
+ kgt kL2 (0,T ;H −0.5 (∂Ω)) + C kuκt (s)k2L2 (Ω) + kDuκ (s)k2L2 (Ω) ds ;
0

thus the Gronwall inequality and (6.26) imply that


Z t
kuκt (t)k2L2 (Ω) + kDuκ (t)k2L2 (Ω) + κ kDuκt (s)k2L2 (Ω) ds
0
h
≤ C ku0 kH 1 (Ω) + ku1 kL2 (Ω) + kf k2L2 (0,T ;L2 (Ω))
2 2
(7.21)
i
+ kgk2L2 (0,T ;H −0.5 (∂Ω)) + kgt k2L2 (0,T ;H −0.5 (∂Ω)) eCt ;

thus uκ is uniformly bounded in L∞ (0, T ; H 1 (Ω)), uκt is uniformly bounded



in L∞ (0, T ; L2 (Ω)), and κDuκt is uniformly bounded in L2 (0, Tκ ; L2 (Ω)).
We also need to establish a uniform bound for uκtt in L2 (0, T ; H 1 (Ω)0 ).
By the variational formulation (7.19) we find that

kuκtt kH 1 (Ω)0 = sup huκtt , ϕiH 1 (Ω) ≤ kf kL2 (Ω) + κkDuκt kL2 (Ω) + kDuκ kL2 (Ω) ;
kϕkH 1 (Ω) =1

thus (7.21) implies that


Z t h
kuκtt (s)k2H 1 (Ω)0 ds ≤ C ku0 k2H 1 (Ω) + ku1 k2L2 (Ω) + kf k2L2 (0,T ;L2 (Ω)) (7.22)
0
i
+ kgk2L2 (0,T ;H 0.5 (∂Ω)) + kgt k2L2 (0,T ;H −0.5 (∂Ω)) .

The combination of (7.21) and (7.22) then gives


h i
max kuκt (t)kL2 (Ω) + kuκ (t)kH 1 (Ω) + kuκtt (t)kL2 (0,T ;H 1 (Ω)0 )
t∈[0,T ]
h
≤ C ku0 kH 1 (Ω) + ku1 kL2 (Ω) + kf kL2 (0,T ;L2 (Ω)) (7.23)
i
+ kgkL2 (0,T ;H 0.5 (∂Ω)) + kgt kL2 (0,T ;H −0.5 (∂Ω))

for some constant C = C(T ).


280 CHAPTER 7. Linear Hyperbolic Equations

A κ-independent time of existence

By the contraction mapping theorem, the solution uκ exists in a time interval


of length at least Tκ . If the solution uκ fails to exist, beyond certain T > Tκ
the function uκ will fail to belong to L2 (0, T ; H 1 (Ω)) which is the same as
saying that kuκ kL2 (0,T ;H 1 (Ω)) = ∞. However, we note that (7.21) holds as
long as the solution exists, so it is impossible that uκ blows up in finite time.
Therefore, the time of existence Tκ can be make independent of κ, and we
set Tκ = T . Moreover, the solution exists uniquely in the time interval [0, T ].

The limit as κ approaches zero



By the uniform bound of κDuκt in L2 (0, T ; L2 (Ω)), for all ϕ ∈ H 1 (Ω) we
have
Z T √
Z T √
κ (Duκt , Dϕ)L2 (Ω) dt = κ ( κDuκt , Dϕ)L2 (Ω) dt → 0 as κ → 0.
0 0
Therefore, integrating (7.19) over the time interval (0, T ) and passing κ to
the limit, we obtain that
Z Th i Z T
f (t), ϕ L2 (Ω) dt ∀ ϕ ∈ H 1 (Ω),
 
hutt (t), ϕiH 1 (Ω) + Du(t), Dϕ L2 (Ω) dt =
0 0
here u is the weak limit of uκ in 2 1
L (0, T ; H (Ω)). This implies that u satisfies
the variational formulation (7.6). Moreover, (7.23) also implies that uκtt is
uniformly bounded in L2 (0, T ; H 1 (Ω)0 ), while Duκ and Duκ are uniformly
bounded in Lp (0, T ; L2 (Ω)). Therefore, the weak limit u satisfies
h
kukLp (0,T ;H 1 (Ω)) + kut kLp (0,T ;L2 (Ω)) + kutt kL2 (0,T ;H 1 (Ω)0 ) ≤ C ku0 kH 1 (Ω)
i
+ ku1 kL2 (Ω) + kf kL2 (0,T ;L2 (Ω)) + kgkL2 (0,T ;H 0.5 (∂Ω)) + kgt kL2 (0,T ;H −0.5 (∂Ω))

for some constant C = C(T ), independent of p. Since lim khkLp (0,T ) =


p→∞
khkL∞ (0,T ) , we conclude that u satisfies
h
kukL∞ (0,T ;H 1 (Ω)) + kut kL∞ (0,T ;L2 (Ω)) + kutt kL2 (0,T ;H 1 (Ω)0 ) ≤ C ku0 kH 1 (Ω)
i
+ ku1 kL2 (Ω) + kf kL2 (0,T ;L2 (Ω)) + kgkL2 (0,T ;H 0.5 (∂Ω)) + kgt kL2 (0,T ;H −0.5 (∂Ω)) ;

so u ∈ L∞ (0, T ; H 1 (Ω)) with time derivatives ut ∈ L∞ (0, T ; L2 (Ω)) and


utt ∈ L2 (0, T ; H 1 (Ω)0 ). Finally,
Z t Z t
uκ (t) = u0 + w(s) ds and w(t) = u1 + uκtt (s) ds
0 0
§7.2 The method of parabolic regularization - the simple case 281

imply that u(0) = u0 and ut (0) = u1 . Therefore, u is the unique weak


solution to (7.14).

7.2.2 Regularity theory with smooth data

Before proceeding, we introduce the following spaces.

1. the space of solutions: the space consists of space-time dependent


functions so that one time derivative scales like one space derivative.
To be more concrete, for k ≥ 2 and k ∈ N,
n o
W (T ; Ω) ≡ u ∈ L (0, T ; H (Ω)) ∂tj u ∈ L∞ (0, T ; H k−j (Ω)) if j ≤ k
k ∞ k

k
k∂tj ukL∞ (0,T ;H k−j (Ω)) ;
X
equipped with norm kukW k (T ;Ω) =
j=0

2. the space of interior forcing functions: for k ≥ 2 and k ∈ N,


n
FΩk−2 (T ) ≡ f ∈ C([0, T ]; H k−2 (Ω)) ∂tk−1f ∈ L2 (0, T ; L2 (Ω)),

o
∂tj f ∈ C([0, T ]; H k−j−2 (Ω)) if j ≤ k − 2

equipped with norm

k−2
∂tj kf kC([0,T ];H k−j−2 (Ω)) + k∂tk−1f kL2 (0,T ;L2 (Ω)) ;
X
kf kW k−2 (T ) =

j=0

3. the space of forcing functions on the boundary: for k ≥ 2 and k ∈ N,


n
k−1.5
G∂Ω (T ) ≡ g ∈ C([0, T ]; H k−1.5 (∂Ω)) ∂tk g ∈ L2 (0, T ; H −0.5 (∂Ω)),

o
∂tj g ∈ C([0, T ]; H k−j−0.5 (∂Ω)) if j ≤ k − 1

equipped with norm

k
k∂tj gkC([0,T ];H k−j−1.5 (∂Ω)) + k∂tkgkL2 (0,T ;H −0.5 (∂Ω)) .
X
kgkG k−1.5 (T ) =
∂Ω
j=0

In order to show that the weak solution u to (7.14) possesses higher reg-
ularity, say u ∈ L2 (0, T ; H m+1 (Ω)), we do not look at the zero-th parabolic
282 CHAPTER 7. Linear Hyperbolic Equations

regularization of (7.14), but instead start from the m-times time-differentiated


version of the problem of (7.15)

wm = ∂tm w = ∂tm+1 uκ in Ω × (0, T ) , (7.24a)


wmt − κ∆wm − ∆wm−1 = ∂tm f in Ω × (0, T ) , (7.24b)
∂wm ∂wm−1
κ + = ∂tm g on ∂Ω × (0, T ) , (7.24c)
∂N ∂N
wm = um+1 , wm−1 = um on Ω × {t = 0} , (7.24d)

along with its variational formulation

 
wmt (t), ϕ H 1 (Ω) + κ Dwm (t), ϕ L2 (Ω) + Dwm−1 (t), ϕ L2 (Ω) (7.25)
= ∂tm f (t), ϕ L2 (Ω) + ∂tm g(t), ϕ H 0.5 (∂Ω) ∀ ϕ ∈ H 1 (Ω), a.e. t ∈ (0, T ) ,


where in (7.24d) um and um+1 are given initial data whose precise forms
will be specified later.
The argument of showing that (7.15) has a unique weak solution can be
used to show that (7.24) has a unique weak solution wm ∈ L2 (0, T ; H 1 (Ω))
with wmt ∈ L2 (0, T ; H 1 (Ω)0 ) satisfying (7.25). Integrating (7.25) in time
over the time interval (0, t), we find that
  Z t   Z t 
wm (t), ϕ L2 (Ω)+ κ D wm (s)ds, Dϕ 2 + D wm−1 (s)ds, Dϕ 2
0 L (Ω) 0 L (Ω)
   
= ∂tm−1f (t)−(∂tm−1f )(0)+ um+1 , ϕ 2 + ∂tm−1g(t)−(∂tm−1g)(0), ϕ 2 ,
L (Ω) L (∂Ω)

Z t
and letting wk = uk+1 + wk+1 (s)ds for k = 1, · · · , m further implies that
0
  
wm−1 (t), ϕ L2 (Ω) + κ Dwm−1 , Dϕ L2 (Ω) + Dwm−2 , Dϕ L2 (Ω)
= ∂tm−1f (t) − (∂tm−1f )(0) + um+1 − κ∆um − ∆um−1 , ϕ L2 (Ω) (7.26)


 ∂um ∂um−1 
+ ∂tm−1g(t) − (∂tm−1g)(0) + κ + ,ϕ 2 .
∂N ∂N L (∂Ω)

In other words, wm−1 is the weak solution to

∂t wm−1 − κ∆wm−1 − ∆wm−2 = Fm−1 in Ω × (0, T ) ,


∂wm ∂wm−1
κ + = Gm−1 on ∂Ω × (0, T ) ,
∂N ∂N
wm−1 = um , wm−2 = um−1 on Ω × {t = 0} ,
§7.2 The method of parabolic regularization - the simple case 283

where

Fm−1 = ∂tm−1 f (t) − (∂tm−1 f )(0) + um+1 − κ∆um − ∆um−1 ,


∂um ∂um−1
Gm−1 = ∂tm−1 g(t) − (∂tm−1 g)(0) + κ + .
∂N ∂N
Integrating (7.26) in time over the time interval (0, t) for another (m − 1)
times, we find that w is the weak solution to

wt − κ∆w− ∆uκ = f + δf in Ω × (0, T ) , (7.27a)


∂w ∂uκ
κ + = g + δg on ∂Ω × (0, T ) , (7.27b)
∂N ∂N
w = u1 , uκ = u0 on Ω × {t = 0}. (7.27c)

where
m−1
X t`  ∂`f

δf = u`+2 − (0) − κ∆u`+1 − ∆u` , (7.28a)
`! ∂t`
`=0
m−1
X t`  t` ∂u`+1 ∂u` ∂ ` g 
δg = κ + − ` (0) . (7.28b)
`! `! ∂N ∂N ∂t
`=0

For simplicity, we first assume that the initial data u0 , u1 , the forcing func-
tion f , g are smooth, and satisfy the compatibility conditions

∂u` ∂`g
= ` (0) on ∂Ω, ` = 0, 1, · · · , m − 1,
∂N ∂t
∂ `−1
where u`+1 = (f + ∆u + κ∆ut ) for ` = 1, · · · , m. Note that then
∂t`−1 t=0
δf = 0, and
m−1
X t` ∂u`+1
δg = κ . (7.28b0 )
`! ∂N
`=0

As discussed above, in order to show that the weak solution u to (7.14)


belonging to L2 (0, T ; H m+1 (Ω)), instead of studying (7.15) we focus on

wt − κ∆w = f + ∆uκ in Ω × (0, T ) , (7.29a)


∂w ∂uκ
κ =g− + δg on ∂Ω × (0, T ) , (7.29b)
∂N ∂N
w = u1 on Ω × {t = 0} , (7.29c)
Z t
uκ = u0 + w(s) ds in × Ω × (0, T ) . (7.29d)
0
284 CHAPTER 7. Linear Hyperbolic Equations

We emphasize again that the boundary condition (7.29b) is designed in


such a way that the compatibility conditions for the parabolic equation
of w are valid up to m-th order. By the smoothness assumptions on the
initial and boundary data, using the fixed-point argument we obtain that
2m (T ; Ω). The key then is to show that ∂ ` u ∈
the unique solution w ∈ VN t
L2 (0, T ; H m+1−` (Ω)) is uniformly bounded for ` = 0, 1, · · · , m + 1. In the
following, we take m = 1 and m = 2 as examples to illustrate how κ-
independent estimates are obtained.

κ-independent estimates, the case m = 1

We study the time-differentiated version of the problem of (7.29):

wtt − κ∆wt − ∆w = ft in Ω × (0, T ) , (7.30a)


∂wt ∂w
κ + = gt on ∂Ω × (0, T ) , (7.30b)
∂N ∂N
wt = f (0) + ∆u0 + κ∆u1 on Ω × {t = 0} . (7.30c)

We note that by our smoothness assumptions of the forcing function, bound-


ary and initial data, Theorem 6.29 implies that the solution wt to (7.30)
2 (T ; Ω).
exists and belongs to VN
For the uniform bound of ∂t` uκ in L2 (0, T ; H 2−` (Ω)) for ` = 1, 2, we test
(7.30a) against wt and find that
1 dh i
kwt k2L2 (Ω) + kDwk2L2 (Ω) + κkDwt k2L2 (Ω)
2 dt
1 1
≤ kft k2L2 (Ω) + kwt k2L2 (Ω) + hgt , wt iH 0.5 (∂Ω) .
2 2
Integrating the inequality above in time over the time interval (0, t), by
estimates similar to (7.20) we find that
Z t
2 2
kwt (t)kL2 (Ω)+ kDw(t)kL2 (Ω)+ κ kDwt (s)k2L2 (Ω) ds
0
h
≤ C kf (0)kL2 (Ω)+ ku0 kH 2 (Ω)+ ku1 k2H 1 (Ω)+ kft k2L2 (0,T ;L2 (Ω))
2 2

i
+ kgt k2L2 (0,T ;H −0.5 (∂Ω))+ kgtt k2L2 (0,T ;H −0.5 (∂Ω))
h
≤ C ku0 k2H 2 (Ω)+ ku1 k2H 1 (Ω)+ kf k2L2 (0,T ;L2 (Ω))+ kft k2L2 (0,T ;L2 (Ω))
(7.31)
i
+ kgt k2L2 (0,T ;H −0.5 (∂Ω))+ kgtt k2L2 (0,T ;H −0.5 (∂Ω)) .
§7.2 The method of parabolic regularization - the simple case 285


Therefore, we establish the uniform boundedness of uκtt , uκt , and κuκtt in
L∞ (0, T ; L2 (Ω)), L∞ (0, T ; H 1 (Ω)), and L2 (0, T ; H 1 (Ω)), respectively.
The last piece of the puzzle is to obtain a uniform bound for uκ ∈
L2 (0, T ; H 2 (Ω)). This relies on an elliptic-type estimate, and we first focus
on the following

Lemma 7.9. Let u0 ∈ H m (Ω) and ut ∈ L2 (0, T ; H m (Ω)) satisfies

κut + u = F

for some F ∈ C([0, T ]; H m (Ω)). Then

max ku(t)kH m (Ω) ≤ ku0 kH m (Ω) + max kF (t)kH m (Ω) . (7.32)


t∈[0,T ] t∈[0,T ]

Proof. For almost all t ∈ [0, T ],

kκut (t) + u(t)k2H m (Ω) = kF (t)k2H m (Ω)

which implies

κ2 kut k2H m (Ω) + 2κ(ut , u)H m (Ω) + kuk2H m (Ω) = kF k2H m (Ω) .

By Lemma 6.9,
d
κ ku(t)k2H m (Ω) + ku(t)k2H m (Ω) ≤ kF (t)k2H m (Ω) . (7.33)
dt
Suppose that

max ku(t)kH m (Ω) = ku(t0 )kH m (Ω) . (7.34)


t∈[0,T ]

If t0 = 0, then max ku(t)kH m (Ω) ≤ ku0 kH m (Ω) . If t0 > 0, we integrate (7.33)


t∈[0,T ]
in time over the time interval (t, t0 ) with t < t0 , and Young’s inequality
implies that
h i Z t0 Z t0
2 2
κ ku(t0 )kH m (Ω) − ku(t)kH m (Ω) + ku(s)k2H m (Ω) ds ≤ kF (s)k2H m (Ω) ds.
t t

Assumption (7.34) then further implies that


Z t0 Z t0
ku(s)k2H m (Ω) ds ≤ kF (s)k2H m (Ω) ds ∀ t ∈ (0, t0 ).
t t
286 CHAPTER 7. Linear Hyperbolic Equations

Dividing through both sides by t0 − t and passing t → t0 , the fundamental


theorem of Calculus suggests that

max ku(t)kH m (Ω) = ku(t0 )kH m (Ω) ≤ kF (t0 )kH m (Ω) ≤ max kF (t)kH m (Ω)
t∈[0,T ] t∈[0,T ]

which concludes the lemma. 

The following corollary is important for the discussion in the rest of the
chapter.

Corollary 7.10. Let ∂Ω be of class C m+1 , and aij ∈ C m (Ω×[0, T ]) satisfies


aij ξj ξi ≥ λ 0 |ξ|2 for some constant λ 0 > 0. Suppose that w = κut + u is a
weak solution to
∂  ij ∂w 
− a =f in Ω × (0, T ) , (7.35a)
∂xi ∂xj
∂w
aij Ni = g on ∂Ω × (0, T ) . (7.35b)
∂xj
for some f and g.

1. If u0 ∈ H m+1 (Ω), f ∈ L2 (0, T ; H m−1 (Ω)) and g ∈ L2 (0, T ; H m−0.5 (∂Ω)),


then u ∈ C([0, T ]; H m+1 (Ω));

2. If in addition f ∈ C([0, T ]; H m−1 (Ω)) and g ∈ C([0, T ]; H m−0.5 (∂Ω)),


then u ∈ C([0, T ]; H m+1 (Ω)) satisfies
h
max ku(t)kH m+1 (Ω) ≤ ku0 kH m+1 (Ω) + C max kf (t)kH m−1 (Ω)
t∈[0,T ] t∈[0,T ]
i
+ kg(t)kH m−0.5 (∂Ω) + ku(t)kL2 (Ω)
(7.36)
for some C = C(kakC m (Ω×[0,T ]) , λ 0 ).

Proof. By the elliptic regularity


h i
kwkH m+1 (Ω) ≤ C kf kH m−1 (Ω) + kgkH m−0.5 (∂Ω) + kwkL2 (Ω) (7.37)

for some C = C(kakC m (Ω×[0,T ]) , λ 0 ). Moreover, solving κut + u = w implies


that
t
1
Z
t−s
− κt
u(t) = e u0 + e− κ w(s) ds ; (7.38)
κ 0

thus u ∈ C([0, T ]; H m+1 (Ω)) and ut = κ−1 (w − u) ∈ L2 (0, T ; H m+1 (Ω)).


(7.36) then follows from Lemma 7.9. 
§7.2 The method of parabolic regularization - the simple case 287

Remark 7.11. For the case m = 1, estimate (7.37) and that u is uni-
formly bounded in H 2 (Ω) imply that κut is uniformly bounded in H 2 (Ω).

On the other hand, estimate (7.31) implies that κut is uniformly bounded

in H 1 (Ω); thus a multiplication of κ gives extra (uniform) spatial regular-
ity.

Now we turn our attention back to the uniform bound of the solution uκ
of (7.29) in L2 (0, T ; H 2 (Ω)). First of all, by (7.31),
Z t
κ
ku (t)kL2 (Ω) ≤ ku0 kL2 (Ω) + kuκt (s)kL2 (Ω) ds
0
Z th Z s i
≤ ku0 kL2 (Ω) + ku1 kL2 (Ω) + kuκtt (s 0 )kL2 (Ω) ds 0 ds
0 0
h
≤ C ku0 kH 2 (Ω) + ku1 kH 1 (Ω) + kft k2L2 (0,T ;L2 (Ω))
2 2
(7.39)
i
+ kgt k2L2 (0,T ;H −0.5 (∂Ω)) + kgtt k2L2 (0,T ;H −0.5 (∂Ω))

for some constant C = C(T ). Since uκ satisfies

−κ∆uκt − ∆uκ = f − uκtt in Ω,


∂uκ ∂uκ ∂u1
κ t + =g+κ on ∂Ω ,
∂N ∂N ∂N
by Corollary 7.10,
h
max kuκ (t)kH 2 (Ω) ≤ ku0 kH 2 (Ω) + C max kf (t) − uκtt (t)kL2 (Ω)
t∈[0,T ] t∈[0,T ]
∂u1 i
+ g(t) + κ + kuκ (t)kL2 (Ω) .

∂N H (∂Ω)
0.5

As a consequence, we conclude from (7.31) and (7.39) that

max kuκ (t)kH 2 (Ω)


t∈[0,T ]
h
≤ ku0 kH 2 (Ω) + C max kf (t)kL2 (Ω) + kuκtt (t)kL2 (Ω) + kg(t)kH 0.5 (∂Ω)
t∈[0,T ]
i
+ κku1 kH 2 (Ω) + kuκ (t)kL2 (Ω)
h i
≤ C ku0 kH 2 (Ω)+ ku1 kH 1 (Ω)+ κku1 kH 2 (Ω)+ kf kW 0 (T )+ kgkG 0.5 (T ) .
Ω ∂Ω
(7.40)

This estimate implies that uκ is uniformly bounded in L∞ (0, T ; H 2 (Ω)),


while Duκt and uκtt are uniformly bounded in L∞ (0, T ; L2 (Ω)) by (7.31);
288 CHAPTER 7. Linear Hyperbolic Equations

thus the weak limit u of uκ in Lp (0, T ; H 2 (Ω)) satisfies

kukL∞ (0,T ;H 2 (Ω)) + kut kL∞ (0,T ;H 1 (Ω)) + kutt kL∞ (0,T ;L2 (Ω))
h i (7.41)
≤ C ku0 kH 2 (Ω) + ku1 kH 1 (Ω) + kf kW 0 (T ) + kgkG 0.5 (T ) .
Ω ∂Ω

∂u0
Moreover, by the compatibility condition = g(0), uκ satisfies the vari-
∂N
ational formulation

uκtt (t), ϕ L2 (Ω) + κ Duκt (t), Dϕ L2 (Ω) + Duκ (t), Dϕ L2 (Ω)


  
(7.42)
∀ ϕ ∈ H 1 (Ω), a.e. t ∈ (0, T ).
 
= f (t), ϕ L2 (Ω) + g(t), ϕ L2 (∂Ω)

Therefore, integrating (7.42) in time over the time interval (0, T ) and pass-
ing κ → 0, we find that there exists a u ∈ L2 (0, T ; H 2 (Ω)) with ut ∈
L2 (0, T ; H 1 (Ω)) and utt ∈ L2 (0, T ; L2 (Ω)) satisfying
Z Th i
 
utt (s), ϕ L2 (Ω) + Du(s), Dϕ L2 (Ω) ds
0
Z Th i
∀ ϕ ∈ H 1 (Ω)
 
= f (s), ϕ L2 (Ω) + g(s), ϕ L2 (∂Ω) ds
0

Remark 7.5 implies u satisfies


 
utt (t), ϕ L2 (Ω) + Du(t), Dϕ L2 (Ω)
∀ ϕ ∈ H 1 (Ω), a.e. t ∈ (0, T ) .
 
= f (t), ϕ L2 (Ω) + g(t), ϕ L2 (∂Ω)

Moreover, taking the limit of


Z t Z t
κ
u (t) = u0 + uκt (s) ds, uκt (t) = u1 + uκtt (s) ds,
0 0

we conclude that u(0) = u0 and ut (0) = u1 . Thus we prove the weak solution
u to (7.14) indeed possesses better regularity and satisfies (7.41), provided
that the initial data and the forcing functions are smooth, as well as the
first-order compatibility condition.

κ-independent estimates, the case m = 2

For the case m = 2, we assume that first-order and second-order compati-


bility conditions
∂u0 ∂u1
= g(0), = gt (0) on ∂Ω.
∂N ∂N
§7.2 The method of parabolic regularization - the simple case 289

Note that the compatibility conditions imply that

kg(0)kH 1.5 (∂Ω) ≤ Cku0 kH 3 (Ω) , kgt (0)kH 0.5 (∂Ω) ≤ Cku1 kH 2 (Ω) .

In order to obtain uκ ∈ L∞ (0, T ; H 3 (Ω)) with κ-independent estimates, we


need to show that wt ∈ L∞ (0, T ; H 1 (Ω)). It requires the study of

wttt − κ∆wtt − ∆wt = ftt in Ω × (0, T ) , (7.43a)


∂wtt ∂wt
κ + = gtt on ∂Ω × (0, T ) , (7.43b)
∂N ∂N
wt = u2 , wtt = u3 on Ω × {t = 0} , (7.43c)

which is the second time-differentiated version of the problem of

wt − κ∆w = f + ∆uκ in Ω × (0, T ) , (7.44a)


∂w ∂uκ ∂u1 ∂u2
κ + =g+κ + κt on ∂Ω × (0, T ) , (7.44b)
∂N ∂N ∂N ∂N
w = u1 on Ω × {t = 0} , (7.44c)
Z t
κ
u (t) = u0 + w(s) ds, (7.44d)
0

where in (7.43c) and (7.44b), u2 ≡ f (0)+∆u0 +κ∆u1 and u3 = ft (0)+∆u1 +


κ∆u2 . The weak solution wtt to (7.43) belongs to L2 (0, T ; H 1 (Ω)) with time
derivative wttt ∈ L2 (0, T ; H 1 (Ω)0 ), and similar to (7.31), w satisfies
h i Z T
sup kwtt (t)k2L2 (Ω)+ kDwt (t)k2L2 (Ω) +κ kDwtt (s)k2L2 (Ω) ds
t∈[0,T ] 0
h
≤ C ku3 k2L2 (Ω)+ ku2 k2H 1 (Ω)+ kftt k2L2 (0,T ;L2 (Ω))+ kgtt k2L2 (0,T ;H −0.5 (∂Ω))
i
+ kgttt k2L2 (0,T ;H −0.5 (∂Ω))

for some constant C = C(T ). By the second-order compatibility condition


∂u1
= gt (0) on ∂Ω,
∂N
wt satisfies (the variational formulation of)

wtt − κ∆wt − ∆w = ft in Ω × (0, T ) , (7.45a)


∂wt ∂w ∂u1
κ + = gt + κ on ∂Ω × (0, T ) , (7.45b)
∂N ∂N ∂N
w = u1 , wt = u2 on Ω × {t = 0}; (7.45c)
290 CHAPTER 7. Linear Hyperbolic Equations

thus by Corollary 7.10,


h
max kw(t)kH 2 (Ω) ≤ ku1 kH 2 (Ω) + C max kft (t) − wtt (t)kL2 (Ω)
t∈[0,T ] t∈[0,T ]
∂u1 i
+ gt (t) + κ + kw(t)k

2
L (Ω)
∂N H 0.5 (∂Ω)

h
≤ C ku3 kL2 (Ω) + ku2 kH 1 (Ω) + (1 + κ)ku1 kH 2 (Ω)
+ kft kL2 (0,T ;L2 (Ω)) + kftt kL2 (0,T ;L2 (Ω)) + kgt kC([0,T ];H 0.5 (∂Ω))
i
+ kgtt kL2 (0,T ;H −0.5 (∂Ω)) + kgttt kL2 (0,T ;H −0.5 (∂Ω))

for some constant C = C(T ). By the first-order compatibility condition


∂u0
= g(0) on ∂Ω,
∂N
w satisfies (7.44); thus by Corollary 7.10 again,

max kuκ (t)kH 3 (Ω)


t∈[0,T ]
h 
≤ ku0 kH 3 (Ω) + C ku3 kL2 (Ω) + ku2 kH 1 (Ω) + κ ku1 kH 3 (Ω) + ku2 kH 3 (Ω)
+ kf kC([0,T ];H 1 (Ω)) + kftt kL2 (0,T ;L2 (Ω)) + kgkC([0,T ];H 1.5 (∂Ω))
i
+ kgtt kL2 (0,T ;H −0.5 (∂Ω)) + kgttt kL2 (0,T ;H −0.5 (∂Ω)) .

Since u2 = f (0) + ∆u0 + κ∆u1 and u3 = ft (0) + ∆u1 + κ∆u2 ,

ku3 kL2 (Ω) + ku2 kH 1 (Ω) + κku2 kH 3 (Ω)


≤ kf (0)kH 1 (Ω) + kft (0)kL2 (Ω) + ku0 kH 3 (Ω) + ku1 kH 2 (Ω)
+ κku1 kH 3 (Ω) + κkf (0)kH 3 (Ω) + κku0 kH 5 (Ω) + κ2 ku1 kH 5 (Ω) ,

from which we conclude that


h i
max kuκ (t)kH 3 (Ω)+ kuκt (t)kH 2 (Ω)+ kuκtt (t)kH 1 (Ω)+ kuκttt (t)kL2 (Ω)
t∈[0,T ]
h
≤ C ku0 kH 3 (Ω)+ ku1 kH 2 (Ω)+ κku0 kH 5 (Ω)+ κku1 kH 3 (Ω) (7.46)
i
+ κkf (0)kH 3 (Ω)+ κ2 ku1 kH 5 (Ω)+ kf kF 1 (T ) + kgkG 1.5 (T )
Ω ∂Ω

for some constant C = C(T ). (7.46) further implies that the weak limit u
of uκ satisfies
h i
kukW 3 (T ;Ω) ≤ C ku0 kH 3 (Ω) + ku1 kH 2 (Ω) + kf kW 1 (T ) + kgkG 1.5 (T ) . (7.47)
Ω ∂Ω
§7.2 The method of parabolic regularization - the simple case 291

We remark here that even though (7.47) does not depend on higher order
norm of u0 , u1 , and f (0), it is required that u0 , u1 , and f (0) to be more
regularity in order to obtain this estimate.

7.2.3 Regularity theory with data possessing limited regu-


larity1
The case m = 1

We first assume that u0 ∈ H 2 (Ω), u1 ∈ H 1 (Ω), f ∈ FΩ0 (T ), and g ∈ G∂Ω


0.5 (T )

as well as the first-order compatibility condition

∂u0
= g(0) on ∂Ω.
∂N

Note that in (7.40), u1 is required to belong to H 2 (Ω) due to the extra


∂u1
term κ that we add into the boundary condition in order to make sure
∂N
the validity of the compatibility condition for the parabolic equation of w.
However, if we can regularize u1 by the parameter κ in such a way that

κku1κ kH 2 (Ω) ≤ Cku1 kH 1 (Ω) and u1κ → u1 as κ → 0, (7.48)

where u1κ denotes some smooth version of u1 , then by considering

wt − κ∆w = f + ∆uκ in Ω × (0, T ) , (7.49a)


∂w ∂uκ ∂u1κ
κ =g− +κ on ∂Ω × (0, T ) , (7.49b)
∂N ∂N ∂N
w = u1κ on Ω × {t = 0} , (7.49c)
Z t
uκ = u0 + w(s) ds in × Ω × (0, T ) , (7.49d)
0

the same procedure implies that uκ satisfies (7.40) with u1 replaced by u1κ ;
thus by (7.48) the weak limit u satisfies (7.41). It remains to show that
(7.48) is valid for some u1κ , and ut (0) = u1 for the weak limit u.
There are several ways of regularizing u1 so that (7.48) holds. The easiest
way should be done by the mollification. Let Ek : H k (Ω) → H k (Rn ) denote
an extension operator satisfying

1. Ek u ∈ H k (Rn ), and Ek u is compactly supported;


1
The reader may skip this section on the first reading
292 CHAPTER 7. Linear Hyperbolic Equations

2. Ek u = u in Ω, and

3. Ek u satisfies kEk ukH k (Rn ) ≤ CkukH k (Ω) for all u ∈ H k (Ω).

We define u1κ as ηκ ∗ (E1 u1 ), where ηκ is the usual mollifiers defined in


Section ??. By the property of convolutions,

ku1κ kH k+` (Ω) ≤ Cκ−` ku1 kH k (Ω) ;

thus (7.48) holds.


Finally, by the fundamental theorem of Calculus,
Z t
κ
ut (t) = u1κ + uκtt (s)ds ;
0

thus passing κ to 0, by u1κ → u1 in H 1 (Ω), uκt * ut in L2 (0, T ; H 1 (Ω)), and


uκtt * utt in L2 (0, T ; L2 (Ω)), we conclude that
Z t
ut (t) = u1 + utt (s)ds
0

which implies that ut (0) = u1 .


We also emphasize that in the current discussion we do not modify u0 , so
the first-order compatibility condition of the parabolic equation of w remains
valid. The next discussion will focus on the case that we have to regularize
u0 as well.

The case that m = 2

Similar to the discussion in the previous case, the presence of κku0 kH 5 (Ω) ,
κku1 kH 3 (Ω) , κkf (0)kH 3 (Ω) and κ2 ku1 kH 5 (Ω) in estimate (7.46) implies that
we might need to regularize u0 , u1 , f , as well as g (the reason for the
need of regularizing g can be found in Remark 7.14) in order to conclude
u ∈ L∞ (0, T ; H 3 (Ω)). However, the smooth versions u0κ and u1κ of u0 and
u1 in general do not satisfy
∂u0κ ∂u1κ
= gκ (0), = gκt (0) on ∂Ω
∂N ∂N
which invalidate the compatibility conditions for the parabolic equation
(7.49). As a consequence, integrating the variational formulation of (7.43)
(with u2 and u3 replaced by u2κ and u3κ , while the definition of u2κ and u3κ
§7.2 The method of parabolic regularization - the simple case 293

will be stated later) in time will lead to the conclusion that wt is the weak
solution to

wtt − κ∆wt − ∆w = fκt + (δfκ )t in Ω × (0, T ) , (7.50a)


∂wt ∂w
κ + = gκt + (δgκ )t on ∂Ω × (0, T ) , (7.50b)
∂N ∂N
w = u1κ , wt = u2κ on Ω × {t = 0} , (7.50c)

and w is the weak solution to

wt − κ∆w − ∆uκ = fκ + δfκ in Ω × (0, T ) , (7.51a)


∂w ∂uκ
κ + = gκ + δgκ on ∂Ω × (0, T ) , (7.51b)
∂N ∂N
w = u1κ on Ω × {t = 0} , (7.51c)
Z t
κ
u (t) = u0κ + w(s) ds, (7.51d)
0

where
 
δfκ = u2κ − fκ (0) − ∆u0κ − κ∆u1κ + t u3κ − fκt (0) − ∆u1κ − κ∆u2κ ,
∂u0κ ∂u0κ h ∂u
1κ ∂u1κ i
δgκ = κ + − gκ (0) + t κ + − gκt (0) ,
∂N ∂N ∂N ∂N
and u2κ = wt (0) and u3κ = wtt (0) are the initial data for wt and wtt deter-
mined later.
By Corollary 7.10, (7.50) implies that
h
max kw(t)kH 2 (Ω) ≤ ku1κ kH 2 (Ω) + C max kfκt (t) + (δfκ )t − wtt (t)kL2 (Ω)
t∈[0,T ] t∈[0,T ]
i
+ kgκt (t) + (δgκ )t kH 0.5 (∂Ω) + kw(t)kL2 (Ω)
h
≤ C ku3κ kL2 (Ω) + ku2κ kH 1 (Ω) + (1 + κ)ku1κ kH 2 (Ω)
+ ku3κ − fκt (0) − ∆u1κ − κ∆u2κ kL2 (Ω) (7.52)
+ kft kL2 (0,T ;L2 (Ω)) + kftt kL2 (0,T ;L2 (Ω)) + kgt kC([0,T ];H 0.5 (∂Ω))
i
+ kgtt kL2 (0,T ;H −0.5 (∂Ω)) + kgttt kL2 (0,T ;H −0.5 (∂Ω)) ,

here we make use of the estimate


∂u

− gκt (0) 0.5 ≤ Cku1κ kH 2 (Ω)

∂N

H (∂Ω)

to conclude (7.52). We note that here we do not make the substitution u3κ =
fκt (0) − ∆u1κ − κ∆u2κ yet to get rid of ku3κ − fκt (0) − ∆u1κ − κ∆u2κ kL2 (Ω)
294 CHAPTER 7. Linear Hyperbolic Equations

in the estimate. We keep it here for the moment to illustrate one particular
issue that we will encounter later.
Similarly, by Corollary 7.10 again, (7.51) implies that
h
max kuκ (t)kH 3 (Ω) ≤ ku0κ kH 3 (Ω)+ C max kfκ (t)+ δfκ (t)− wt (t)kH 1 (Ω)
t∈[0,T ] t∈[0,T ]
i
+ kgκ (t)+ δgκ (t)kH 1.5 (∂Ω)+ kuκ (t)kL2 (Ω)
h 
≤ ku0κ kH 3 (Ω)+ C ku3κ kL2 (Ω)+ ku2κ kH 1 (Ω)+ κ ku1κ kH 3 (Ω)+ ku2κ kH 3 (Ω)
∂u

+ ku3κ − fκt (0)− ∆u1κ − κ∆u2κ kH 1 (Ω)+ − gκt (0) 1.5

∂N H (∂Ω)
(7.53)
+ kfκ kC([0,T ];H 1 (Ω))+ kfκtt kL2 (0,T ;L2 (Ω))+ kgκ kC([0,T ];H 1.5 (∂Ω))
i
+ kgκtt kL2 (0,T ;H −0.5 (∂Ω))+ kgκttt kL2 (0,T ;H −0.5 (∂Ω)) ,

here we make use of the estimate


∂u

ku2κ − fκ (0)− ∆u0κ − κ∆u1κ kH 1 (Ω) + − gκ (0) 1.5

h ∂N H (∂Ω) i
≤ C ku0κ kH 3 (Ω) + ku2κ kH 1 (Ω) + κku1κ kH 3 (Ω) + kfκ (0)kH 1 (Ω)
(7.54)
to conclude (7.53). We emphasize that we do not make the substitution
∂u0κ
u2κ = fκ (0) + ∆u0κ + κ∆u1κ and that = gκ (0) to conclude (7.54).
∂N
To obtain a κ-independent estimate for uκ ∈ L∞ (0, T ; H 3 (Ω)), we need

1. u3κ − fκt (0) − ∆u1κ − κ∆u2κ uniformly bounded in H 1 (Ω);


∂u1κ
2. − gκt (0) uniformly bounded in H 1.5 (∂Ω), and
∂N

3. κ ku1κ kH 3 (Ω) + ku2κ kH 3 (Ω) is uniformly bounded.

An easy way to achieve the first two requirements is to let u3κ = fκt (0) −
∂u1κ
∆u1κ − κ∆u2κ in Ω and = gκt (0) on ∂Ω. We note that the choice of
∂N
u2κ could be made quite arbitrarily as long as

u2κ → u2 ≡ f (0) + ∆u0 in H 1 (Ω) and κku2κ kH 3 (Ω) ≤ Cku2 kH 1 (Ω) .

After all the analysis above, we state how u0κ , u1κ , u2κ , fκ and gκ
are defined in order to obtain κ-independent estimates. With {ζm }K
m=1
and {θm }K
m=1 denoting the partition of unity and local charts defined in
§7.2 The method of parabolic regularization - the simple case 295

the proof of elliptic regularity, and Λ denoting the horizontal convolution


operator defined in Section 5.8, we define u0κ , fκ , and gκ by

u2κ = ηκ1/3 ∗ (E1 u2 ), fκ = ηκ1/3 ∗ (E1 f ),


K p
X  p i −1
gκ = ζm Λκ1/3 ∗ ( ζm g) ◦ θm ◦ θm .
m=1

Then by the property of the convolution,


h i
κkfκ (0)kH 3 (Ω) ≤ C ku0 kH 3 (Ω) + kf (0)kH 1 (Ω) ,
kgκt (0)kH 1.5 (∂Ω) ≤ Cκ−1/3 kgt (0)kH 0.5 (∂Ω) ≤ Cκ−1/3 ku1 kH 2 (Ω) ,

and as κ → 0,

fκ → f in FΩ1 (T ), 1.5
gκ → g in G∂Ω (T ).

It remains to construct u0κ and u1κ . This can be done by the following

Lemma 7.12. Let f ∈ H k (Ω) and g ∈ H k−1.5 (∂Ω) for some k ≥ 1, and
∂f
= g on ∂Ω. Suppose that fn and gn are sequence of functions such that
∂N
fn → f in H k (Ω), gn → g in H k−1.5 (∂Ω), and

kfn kH k+` (Ω) ≤ K1 (n, `)kf kH k (Ω) , kgn kH k+`−0.5 (∂Ω) ≤ K2 (n, `)kgkH k−1.5 (∂Ω) .

Then the solution un to

un − λ n ∆un = fn in Ω, (7.55a)
∂un
= gn on ∂Ω, (7.55b)
∂N

belongs to H k+` (Ω), and


p
kun kH k+` (Ω) + λ n kun kH k+`+1 (Ω)
q h p i (7.56)
≤ C K21 + K22 kf kH k (Ω) + λ n kgkH k−1.5 (∂Ω) .

Suppose further that λ n > 0 and λ n → 0 as n → 0. Then

1. If for some ` ≥ 0, K1 and K2 are constants independent of n, then un


converges weakly to f in H k (Ω).
296 CHAPTER 7. Linear Hyperbolic Equations

2. un converges (strongly) to f in H k (Ω) as n → ∞, provided that

λ n kun kH k+2 (Ω) → 0 as n → ∞. (7.57)

Proof. By the elliptic regularity,


h i
kun k2H k+` (Ω) + λ n kun k2H k+`+1 (Ω) ≤ C kfn k2H k+` (Ω) + λ n kgn k2H k+`−0.5 (∂Ω)
h i
≤ C K21 kf k2H k (Ω) + λ n K22 kgk2H k−1.5 (∂Ω)
h i
≤ C K21 + K22 kf k2H k (Ω) + λ n kgk2H k−1.5 (∂Ω) .

This proves (7.56).


Now suppose that for some ` ≥ 0, K1 and K2 are constants independent
of n. Then (7.56) implies that un is uniformly bounded in H k+` (Ω). In
particular, un is uniformly bounded in H k (Ω). Therefore, there exists a
subsequence nj of n and u ∈ H k (Ω) such that

unj * u in H k (Ω).

However, (7.56) also implies that λ n un is uniformly bounded in H 1 (Ω).
Passing j to the limit in the variational formulation

(unj , ϕ)L2 (Ω) + λ nj (Dunj , Dϕ)L2 (Ω)


= (fnj , ϕ)L2 (Ω) + λ nj (gnj , ϕ)L2 (∂Ω) ∀ ϕ ∈ H 1 (Ω),

we find that that (u, ϕ)L2 (Ω) = (f, ϕ)L2 (Ω) for all ϕ ∈ H 1 (Ω). This shows
u = f . This procedure also shows that f is the only weak limit of the whole
sequence un ; thus un converges weakly to f in H k (Ω).
If λ n kun kH k+2 (Ω) → 0 as n → ∞, then un − fn → 0 in H k (Ω) as n → ∞
since un − fn = λ n ∆un . 

With Lemma 7.12, we then define u0κ as the solution to

(1 − κ∆)u0κ = ηκ1/3 ∗ (E3 u0 ) in Ω,


∂u0κ
= gκ (0) on ∂Ω .
∂N
and define u1κ as the solution to

(1 − κ∆)u1κ = ηκ1/3 ∗ (E2 u1 ) in Ω,


∂u1κ
= gκt (0) on ∂Ω .
∂N
§7.2 The method of parabolic regularization - the simple case 297

Then by Lemma 7.12 (with k = 2 and ` = 0),


√ h √ i
ku1κ kH 2 (Ω) + κku1κ kH 3 (Ω) ≤ C kηκ1/3 ∗ (E2 u1 )kH 2 (Ω) + κkgκt (0)kH 0.5 (∂Ω)

≤ C(1 + κ)ku1 kH 2 (Ω) ,

Moreover, as κ → 0,

κkηκ1/3 ∗ (E3 u0 )kH 4 (Ω) ≤ Cκ1/6 ku0 kH 3 (Ω) → 0,

κkηκ1/3 ∗ (E2 u1 )kH 3 (Ω) ≤ Cκ1/6 ku1 kH 2 (Ω) → 0;

hence u0κ → u0 in H 3 (Ω) and u1κ → u1 in H 2 (Ω). Therefore, the study of

wt − κ∆w = fκ + ∆uκ in Ω × (0, T ) ,


∂w ∂uκ ∂u1κ ∂u2κ
κ + = gκ + κ + κt on ∂Ω × (0, T ) ,
∂N ∂N ∂N ∂N
j−1
∂t w = ujκ for j = 1, 2, 3 on Ω × {t = 0} ,
Z t
κ
u (t) = u0κ + w(s) ds,
0
implies that
h i
max kuκ (t)kH 3 (Ω)+ kuκt (t)kH 2 (Ω)+ kuκtt (t)kH 1 (Ω)+ kuκttt (t)kL2 (Ω)
t∈[0,T ]
h i
≤ C ku0 kH 3 (Ω) + (1 + κ)ku1 kH 2 (Ω) + kf kW 1 (T ) + kgkG 1.5 (T ) ;
Ω ∂Ω

thus the weak limit u of uκsatisfies


h i
kukW 3 (T ;Ω) ≤ C ku0 kH 3 (Ω) + ku1 kH 2 (Ω) + kf kW 1 (T ) + kgkG 1.5 (T ) . (7.58)
Ω ∂Ω

Remark 7.13. As discussed, we can always conclude (7.52) no matter


whether the condition
∂u0κ
= gκ (0) on ∂Ω (7.59)
∂N
is satisfied or not. Nevertheless, we will still regularize u0 in the fashion
that (7.59) is satisfied for consistency.

Remark 7.14. Without regularizing g, by the second-order compatibility


∂u

condition the term − gt (0) in estimate (7.53) is the same as

∂N H 1.5 (∂Ω)
∂(u − u )
1κ 1
. Since u1 ∈ H 2 (Ω) but u1 6∈ H 3 (Ω), it is impossible

∂N
1.5
H (∂Ω)
that on one hand κku1κ kH 3 (Ω) is uniformly bounded (which requires that
∂(u − u )
1κ 1
u1κ ∈ H 3 (Ω)), while on the other hand is bounded

∂N
1.5
H (∂Ω)
(which implies that u1κ at best belongs to H 0.5 (∂Ω)).
298 CHAPTER 7. Linear Hyperbolic Equations

The case that m > 2

Suppose that u0 ∈ H m+1 , u1 ∈ H m , f ∈ FΩm−1 (T ) and g ∈ G∂Ω


m−1.5
(T )
satisfy the first m-th order compatibility conditions
∂u`
= (∂t` g)(0) on ∂Ω, (7.60)
∂N
where u`+2 = (∂t` f )(0)+∆u` for ` = 0, 1, · · · , m−1. Then for 0 ≤ ` ≤ m+1,
h i
ku` kH m−` (Ω) ≤ C ku0 kH m (Ω) + ku1 kH m−1 (Ω) + kf kW m−1 (T ) . (7.61)

We regularize the forcing functions f and g using the mollifier as before:


let fκ and gκ be defined by

fκ = η ∗ (Ef ), (7.62a)
K
X p  p i −1
gκ = ζm Λ ∗ ( ζm g) ◦ θm ◦ θm (7.62b)
m=1

for some  = (κ) < 1 to be determined later. Here we assume (κ) = κγ


for some γ > 0, and the goal is to find γ > 0 small enough so that we can
obtain κ-independent estimates for uκ . In general, as long as (κ) converges
slowly enough to 0 as κ → 0, then κ-independent estimates can be obtained.
As in the previous case, let u0κ be the solution to

u0κ − κ∆u0κ = η ∗ (Em+1 u0 ) in Ω,


∂u0κ
= gκ (0) on ∂Ω.
∂N
Then by elliptic regularity,

ku0κ kH m+`+1 (Ω) + κku0κ kH m+`+2 (Ω)
h √ i
≤ C kη ∗ (Em+1 u0 )kH m+`+1 (Ω) + κkgκ (0)kH m+`+0.5 (∂Ω)
h √ i
≤ C −` ku0 kH m+1 (Ω) + κ−`−1 kg(0)kH m−0.5 (∂Ω)
≤ C−` ku0 kH m+1 (Ω) (7.63)

as long as γ ≤ 1/2 (and from now on we assume that γ ≤ 1/2). Similarly,


let u1κ be the solution to

u1κ − κ∆u1κ = η ∗ (Em u1 ) in Ω,


∂u1κ
= gκt (0) on ∂Ω,
∂N
§7.2 The method of parabolic regularization - the simple case 299

then

ku1κ kH m+` (Ω) + κku1κ kH m+`+1 (Ω) ≤ C−` ku1 kH m (Ω) . (7.64)

For 2 ≤ ` ≤ m − 1, define u`κ as the solution of

u`κ − κ∆u`κ = (∂t`−2 fκ )(0) + ∆u(`−2)κ in Ω, (7.65a)


∂u`κ
= (∂t` gκ )(0) on ∂Ω. (7.65b)
∂N
The elliptic regularity implies that for all k ≥ 0,

ku`κ kH m−`+k+1 (Ω) + κku`κ kH m−`+k+2 (Ω)
h √ i
≤ C k(∂t`−2 fκ )(0) + ∆u(`−2)κ kH m−`+k+1 (Ω) + κk∂t` gκ (0)kH m−`+k+0.5 (∂Ω) .

For ` = 2 and ` = 3, by (7.63) and (7.64) we find that


√ h i
ku2κ kH m+k−1 (Ω) + ku3κ kH m+k−2 (Ω) + κ ku2κ kH m+k (Ω) + ku3κ kH m+k−1 (Ω)
h 3 
X
≤ C−k ku0 kH m+1 (Ω) + ku1 kH m+1 (Ω) + k(∂t`−2 f )(0)kH m−`−1 (Ω)
`=2
i
+ k∂t` g(0)kH m−`−0.5 (∂Ω)
h i
≤ C−k ku0 kH m+1 (Ω) + ku1 kH m+1 (Ω) + kf kW m−1 (T ) ,

which further implies that


√ h i
ku4κ kH m+k−3 (Ω) + ku5κ kH m−k−4 (Ω) + κ ku4κ kH m+k−2 (Ω) + ku5κ kH m+k−3 (Ω)
h 5 
X
≤ C ku2κ kH m+k−1 (Ω) + ku3κ kH m+k−2 (Ω) + k(∂t`−2 fκ )(0)kH m+k−`−1 (Ω)
`=4
i
+ k∂t` gκ (0)kH m+k−`−0.5 (∂Ω)
h i
≤ C−k ku0 kH m+1 (Ω) + ku1 kH m+1 (Ω) + kf kW m−1 (T ) .

By induction we then conclude that for 0 ≤ ` ≤ m − 1 and k ≥ 0,



ku`κ kH m−`+k+1 (Ω) + κku`κ kH m−`+k+2 (Ω)
h i (7.66)
≤ C−k ku0 kH m+1 (Ω) + ku1 kH m (Ω) + kf kW m−1 (T ) .

In other words, we regularize u` in such a fashion that one derivative scales


like a multiplication by −1 , and the smooth version u`κ satisfies (7.65b).
300 CHAPTER 7. Linear Hyperbolic Equations

Finally, for ` = m, m + 1, let

u`κ = (∂t`−2 fκ )(0) + ∆u(`−2)κ + κ∆u(`−1)κ . (7.67)

Then by (7.66), for ` = m, m + 1,


h i
ku`κ kH k (Ω) ≤ C k(∂t`−2 fκ )(0)kL2 (Ω)+ ku(`−2)κ kH k+2 (Ω)+ κku(`−1)κ kH k+2 (Ω)
h i
≤ C−k ku0 kH m+1 (Ω) + ku1 kH m (Ω) + kf kW m−1 (T ) .

In other words, (7.66) also holds for the case ` = m, m + 1.


Having u`κ , fκ and gκ defined, as discussed in the previous section, we
have the following

Definition 7.15 (The parabolic regularization of (7.14)). The m-th order


parabolic regularization of the hyperbolic initial-boundary value problem

utt − ∆u = f in Ω × (0, T ) ,
∂u
=g on ∂Ω × (0, T ) ,
∂N
u = u0 , ut = u1 on Ω × {t = 0} ,

is defined as

wt − κ∆w− ∆u = fκ + δfκ in Ω × (0, T ) , (7.68a)


∂w ∂u
κ + = gκ + δgκ on ∂Ω × (0, T ) , (7.68b)
∂N ∂N
w = u1κ , u = u0κ on Ω × {t = 0} , (7.68c)

where
m−3
X t`  
δfκ = u(`+2)κ − (∂t` fκ )(0) − ∆u`κ − κ∆u(`+1)κ , (7.69a)
`!
`=0
m−1
X t` ∂u(`+1)κ
δgκ = κ . (7.69b)
`! ∂N
`=0

Estimates

In the following, we always assume that 0 < κ  1 (so that   1 as well).


By (7.65a) and (7.66),
m−3
Xh i
k∂tk (δfκ )kH m−k−1 (Ω) ≤ κ k∆u(`+2)κ kH m−k−1 (Ω) + k∆u(`+1)κ kH m−k−1 (Ω)
`=k
h i
k−m+1
≤ Cκ ku0 kH m+1 (Ω) + ku1 kH m (Ω) + kf kW m−1 (T ) . (7.70)

§7.2 The method of parabolic regularization - the simple case 301

Moreover, (7.66) implies that for 0 ≤ k ≤ m,

k∂tk (δgκ )kH m−k−0.5 (∂Ω)


m−1 m−1
X ∂u(`+1)κ
X
≤ Cκ ≤ Cκ ku(`+1)κ kH m−k+1 (Ω)
∂N
m−k−0.5
H (∂Ω)
`=k `=k
m−1
X h i
≤ Cκ k−`−1 ku0 kH m+1 (Ω) + ku1 kH m (Ω) + kf kW m−1 (T )

`=k
h i
≤ Cκk−m ku0 kH m+1 (Ω) + ku1 kH m (Ω) + kf kW m−1 (T ) . (7.71)

Estimates (7.70) and (7.71) suggest that if we let γ = (m + 1)−1 (that is,
(κ) = κ1/(m+1) ), then we have

k∂tk (δfκ )kH m−k−1 (Ω) + k∂tk (δgκ )kH m−k−0.5 (∂Ω)
h i (7.72)
≤ C ku0 kH m+1 (Ω) + ku1 kH m (Ω) + kf kW m−1 (T )

which converges to 0 as κ → 0. Moreover, this choice of  also validate


condition (7.57) in Lemma 7.12; thus u`κ → u` in H m−`+1 (Ω) as κ → 0 for
` = 0, 1, · · · , m − 1. On the other hand,  = κ1/(m+1) implies that

h i
κ k∆u(m−1)κ kH 1 (Ω) + k∆umκ kL2 (Ω)
h i
≤ Cκ−1 ku0 kH m+1 (Ω) + ku1 kH m (Ω) + kf kW m−1 (T ) → 0 as κ → 0;

thus umκ → um in H 1 (Ω) and u(m+1)κ → um+1 in L2 (Ω) as well.


Giving the estimates for the forcing functions δfκ and δgκ , we are now
in the position of performing the estimates for the solution w to (7.68). For
1 ≤ k ≤ m, differentiating (7.68) in time k-times we obtain that

wkt − κ∆wk− ∆wk−1 = ∂tk fκ + ∂tk (δfκ ) in Ω × (0, T ) , (7.73a)


∂wk ∂wk−1
κ + = ∂tk gκ + ∂tk (δgκ ) on ∂Ω × (0, T ) , (7.73b)
∂N ∂N
wk = u(k+1)κ , wk−1 = ukκ on Ω × {t = 0}. (7.73c)

1. The case k = m:
Since ∂tm (δfκ ) = 0 and ∂tm (δgκ ) = 0, testing (7.73a) against wm , similar
302 CHAPTER 7. Linear Hyperbolic Equations

to (7.31) we find that wm satisfies


h i Z T
2 2
sup kwm (t)kL2 (Ω) + kDwm−1 (t)kL2 (Ω) + κ kDwm (s)k2L2 (Ω) ds
t∈[0,T ] 0
h
≤ C ku(m+1)κ k2L2 (Ω) + kumκ k2H 1 (Ω) + k∂tm fκ k2L2 (0,T ;L2 (Ω))
+ k∂tm+1 fκ k2L2 (0,T ;L2 (Ω)) + k∂tm+1 gκ k2L2 (0,T ;H −0.5 (∂Ω))
i
+ k∂tm+2 gκ k2L2 (0,T ;H −0.5 (∂Ω))
h i
≤ C ku0 k2H m (Ω) + ku1 k2H m−1 (Ω) + kf k2W m−1 (T ) + kgk2W m−0.5 (T ) . (7.74)
Ω ∂Ω

By the fundamental theorem of Calculus (7.74) also implies that

max kwm−k (t)kL2 (Ω)


t∈[0,T ]
h i
≤ C ku0 kH m (Ω) + ku1 kH m−1 (Ω) + kf kW m−1 (T ) + kgkW m−0.5 (T )
Ω ∂Ω

for all k = 0, 1, · · · , m.
2. The case 1 ≤ k ≤ m − 1:
Since wkt = wk+1 , by Corollary 7.10 we find that that
h
max kwk−1 (t)kH m−k+1 (Ω) ≤ kukκ kH m−k+1 (Ω) + C max k∂tk fκ (t)kH m−k−1 (Ω)
t∈[0,T ] t∈[0,T ]

+ k∂tk (δfκ )(t)kH m−k−1 (Ω)+ k∂tk gκ (t)kH m−k−0.5 (∂Ω)+ k∂tk (δgκ )(t)kH m−k−0.5 (∂Ω)
i
+ kwk−1 (t)kL2 (Ω)
h i
≤ C ku0 kH m (Ω)+ ku1 kH m−1 (Ω)+ kf kW m−1 (T )+ kgkW m−0.5 (T ) . (7.75)
Ω ∂Ω

3. The case k = 0 or (7.68):


In this case, w0 = uκt , and we conclude from Corollary 7.10 again that
h
max kuκ (t)kH m+1 (Ω) ≤ ku0κ kH m+1 (Ω) + C max kfκ (t) + δfκ (t)kH m−1 (Ω)
t∈[0,T ] t∈[0,T ]
i
κ
+ kgκ (t) + δgκ (t)kH m−0.5 (∂Ω) + ku (t)kL2 (Ω)
h i
≤ C ku0 kH m (Ω) + ku1 kH m−1 (Ω) + kf kW m−1 (T ) + kgkW m−0.5 (T ) . (7.76)
Ω ∂Ω

Combining (7.74), (7.75), and (7.76), we obtain that

kuκ (t)kW m+1 (T ;Ω)


h i
≤ C ku0 kH m (Ω)+ ku1 kH m−1 (Ω)+ kf kW m−1 (T )+ kgkW m−0.5 (T ) .
Ω ∂Ω
7.3. THE PARABOLIC REGULARIZATION OF GENERAL HYPERBOLIC EQUATIONS303

Finally, since u`κ → u` , fκ → f , gκ → g, and δfκ → 0, δgκ → 0, we know


that the weak limit u of uκ is the unique solution to (7.14), and satisfies

kukW m+1 (T ;Ω)


h i
≤ C ku0 kH m+1 (Ω) + ku1 kH m (Ω) + kf kW m−1 (T ) + kgkW m−0.5 (T ) .
Ω ∂Ω
(7.77)

7.3 The parabolic regularization of general hyper-


bolic equations

In the previous section, the parabolic regularization of the wave equation


with Neumann boundary condition is defined. The key is to introduce an ar-
tificial diffusion/viscosity so that the hyperbolic system becomes a parabolic
system. Moreover, if the initial data u0 , u1 or the forcing functions f, g pos-
sess limited regularity, we also need to regularize the data in such a way
that the the compatibility conditions for the new parabolic system (using
the regularized initial data as new initial data and the regularized forcing as
new forcing) are still satisfied. We remark that the compatibility conditions
for the parabolic regularization induces the compatibility conditions for the
hyperbolic equation: since (7.27b) holds for all κ > 0 at t = 0, by (7.28b)
we have
∂u` ∂`g
= ` (0) on ∂Ω, ` = 0, 1, · · · , m − 1.
∂N ∂t
If the coefficients aij is not constant, similar to Definition 6.26 we have
the following

Definition 7.16 (Compatibility conditions). For k ∈ N ∪ {0}, the (k +


1)th-order compatibility condition for the hyperbolic initial-boundary value
problem (7.1a,b,c,d) (resp. (7.1a,b,c,e)) is given by τ uk = 0 on ∂Ω
k−1 
∂uk ∂kg X k ∂ k−` aij ∂u`
  
ij
resp. a (0) Ni = k (0) − Ni k−`
(0) on ∂Ω ,
∂xj ∂t ` ∂t ∂xj
`=0

where u` , ` = 2, · · · , k, is defined by

∂ `
u`+2 = (f + Lu) ` = 0, 1, · · · , k − 2.
∂t` t=0
304 CHAPTER 7. Linear Hyperbolic Equations

Definition 7.16 induces the following

Definition 7.17 (Parabolic regularization). If the coefficients aij , bi , c of


the elliptic operator L, the initial data u0 , u1 , and the forcing functions f ,
g are smooth, then the m-th order parabolic regularization of the hyperbolic
equation

utt + Lu = f in Ω × (0, T ) , (7.78a)


∂u
aij Ni = g on ∂Ω × (0, T ) , (7.78b)
∂xj
u = u0 , ut = u1 on Ω × {t = 0} , (7.78c)

is defined as

Z t
u = u0 + w(s)ds in Ω × (0, T ) , (7.79a)
0
∂  ij ∂w 
wt − κ a = f + Lu in Ω × (0, T ) , (7.79b)
∂xi ∂xj
∂w ∂u
κ aij Ni = g − aij Ni + δg on ∂Ω × (0, T ) , (7.79c)
∂xj ∂xj
w = u1 on Ω × {t = 0} , (7.79d)

where

m−1 `
t` ` ∂uk+1 ∂ `−k
XX  
δg = κNi aij .
`! k ∂xj ∂t`−k t=0
`=0 k=0

When the coefficient aij , bi , c, initial data u0 , u1 and the forcing functions
f, g posses limited regularity, as discussed in Section 7.2.3, a regularization
of the data is required. We regularize the forcing functions f and g as in
(7.62), and regularize the data u` , for 0 ≤ ` ≤ m − 1, by elliptic equations
so that u`κ is the solution to

∂  ij ∂u`κ 
u`κ − κ a (0) = F`κ in Ω, (7.80a)
∂xi κ ∂xj
∂u`κ
aij
κ (0) Ni = G`κ on ∂Ω, (7.80b)
∂xj
7.3. THE PARABOLIC REGULARIZATION OF GENERAL HYPERBOLIC EQUATIONS305

where aij ij
κ is a smooth version of a , G0κ = gκ (0), and

F`κ = η ∗ (Em+1−` u` ) if ` = 0, 1,
F`κ = ∂t`−2 t=0 (fκ + Lκ uκ )

if 2 ≤ ` ≤ m − 1,
`−1   `−r ij
X ` ∂ aκ ∂urκ
G`κ = (∂t` gκ )(0) − Ni (0) if 1 ≤ ` ≤ m − 1,
r ∂t`−r ∂xj
r=0

here with biκ , cκ denoting smooth versions of bi and c, respectively, the elliptic
operator Lκ is defined by

∂  ij ∂v  ∂v
Lκ v ≡ − a + biκ + cκ v.
∂xi κ ∂xj ∂xi

We note that aij i


κ , bκ and cκ can be obtained by mollification (using parameter

 = (κ)) as well. The reason for regularizing the coefficients is to make sure
that u`κ is smooth; for otherwise if aij ∈ C m (Ω × [0, T ]), no matter how
smooth the forcing functions are, u`κ at best belongs to H m+1 (Ω).
Similar to (7.67), for ` = m, m + 1 we define

u`κ = (∂t`−2 fκ )(0) − ∂t`−2 Lκ uκ − κ(aij κ


 
κ ut ,j ),i . (7.81a)

Definition 7.18 (Parabolic regularization). Let u`κ be defined by (7.80) and


(7.81). The m-th order parabolic regularization of the hyperbolic problem

utt + Lu = f in Ω × (0, T ) ,
∂u
aij Ni = g on ∂Ω × (0, T ) ,
∂xj
u = u0 , ut = u1 on Ω × {t = 0} ,

is defined as
Z t
κ
u = u0κ + w(s)ds in Ω × (0, T ) , (7.82a)
0
∂  ij ∂w 
wt − κ a = fκ − Lκ uκ + δfκ in Ω × (0, T ) , (7.82b)
∂xi κ ∂xj
∂w ∂uκ
κ aij
κ Ni = gκ − aij
κ Ni + δgκ on ∂Ω × (0, T ) , (7.82c)
∂xj ∂xj
w = u1κ on Ω × {t = 0}. (7.82d)
306 CHAPTER 7. Linear Hyperbolic Equations

where
m−2
X t`  
u(`+2)κ − ∂t` t=0 fκ − Lκ uκ + κ(aij κ

δfκ = u ,
κ t j i ), ,
`!
`=0
m−1 `
t` ` ∂u(r+1)κ ∂ `−r
XX  
δgκ = κNi aij .
`! r ∂xj ∂t`−r t=0 κ
`=0 r=0

7.3.1 Estimates for the regularized quantities

By the property of convolutions, we assume that

−k ij
1. kaij
κ kC m+k (Ω×[0,T ]) ≤ C ka kC m (Ω×[0,T ]) ;

2. kbiκ kC m+k (Ω×[0,T ]) ≤ C−k kbi kC m (Ω×[0,T ]) ;

3. kcκ kC m+k (Ω×[0,T ]) ≤ C−k kckC m (Ω×[0,T ]) ;

4. kfκ kW m+k−1 (T ) ≤ C−k kf kW m−1 (T ) ;


Ω Ω

5. kgκ kG m+k−0.5 (T ) ≤ C−k kgkW m−0.5 (T ) .


∂Ω ∂Ω

To simplify the notation, throughout the rest of the section we let

N (u0 , u1 , f, g) = ku0 kH m+1 (Ω) + ku1 kH m (Ω) + kf kW m−1 (T ) + kgkW m−0.5 (T ) .


Ω ∂Ω

Estimates for u`κ

The estimates for u`κ is much more complicated than the case that aij = δij ,
so we derive the estimates in great detail. As before, we always assume that
0 < κ  1 and  = (κ) = κγ for some γ < 0.5 to be determined later.

Lemma 7.19. Suppose that ∂Ω is smooth, and aij i


κ , bκ , cκ , fκ and gκ are

defined by mollification, and satisfy the regularity assumptions 1-5. Let u`κ
be defined by (7.80) and (7.81). Then for 0 ≤ ` ≤ m + 1 and 0 ≤ k ≤ Km,

ku`κ kH m−`+k+1 (Ω) + κku`κ kH m−`+k+2 (Ω) ≤ C−k ×
h i
× ku0 kH m+1 (Ω) + ku1 kH m (Ω) + kf kW m−1 (T ) + kgkW m−0.5 (T ) ,
Ω ∂Ω
(7.83)
here the constant C = C kakC m (Ω×[0,T ]) , λ 0 , K).
7.3. THE PARABOLIC REGULARIZATION OF GENERAL HYPERBOLIC EQUATIONS307

Proof. Let 0 ≤ ` ≤ m be fixed. Suppose that we can show that for k ≥ 0,

kF`κ kH m−`+k+1 (Ω) ≤ C−k N (u0 , u1 , f, g) (7.84)

and

κkG`κ kH m−`+k+0.5 (∂Ω) ≤ C−k N (u0 , u1 , f, g), (7.85)

then by elliptic regularity, for k ≥ 0,



ku`κ kH m−`+k+1 (Ω) + κku`κ kH m−`+k+2 (Ω)
m−`+k+1
√ X
≤C κ kaκ (0)kC s (Ω) ku`κ kH m−`+k−s+2 (Ω)
s=1
h √ i
+ C kF`κ kH m−`+k+1 (Ω) + κkG`κ kH m−`+k+0.5 (∂Ω)

√ m−`+k+1
X
≤C κ kaκ (0)kC s (Ω) ku`κ kH m−`+k−s+2 (Ω) (7.86)
s=1
+ C−k N (u0 , u1 , f, g)

for some C = C kaij



κ (0)kC m (Ω) , λ 0 , K .

1. If 0 ≤ k ≤ ` − 1, then kaκ (0)kC s (Ω) ≤ CkakC m (Ω×[0,T ]) . Therefore,


m−`+k+1
X
kaκ (0)kC s (Ω) ku`κ kH m−`+k−s+2 (Ω) ≤ Cmku`κ kH m−`+k+1 (Ω)
s=1

for some constant C = C kakC m (Ω×[0,T ]) . By κ  1, (7.84) and (7.85)
imply that

κku`κ kH m−`+k+2 (Ω) ≤ C−k N (u0 , u1 , f, g),
ku`κ kH m−`+k+1 (Ω) +
(7.87)

where C = C kakC m (Ω×[0,T ]) , λ 0 .

2. If ` ≤ k ≤ ` − 1 + m, then
m−`+k+1
X
kaκ (0)kC s (Ω) ku`κ kH m−`+k−s+2 (Ω)
s=1
m
X m−`+k+1
X 
= + kaκ (0)kC s (Ω) ku`κ kH m−`+k−s+2 (Ω)
s=1 s=m+1
m−`+k+1
X
≤ Cmku`κ kH m−`+k+1 (Ω) + C m−s ku`κ kH m−`+k−s+2 (Ω) .
s=m+1
308 CHAPTER 7. Linear Hyperbolic Equations

By κ  1, we find that


ku`κ kH m−`+k+1 (Ω) + κku`κ kH m−`+k+2 (Ω)
√ m−`+k+1
X
≤C κ m−s ku`κ kH m−`+k−s+2 (Ω) + C−k N (u0 , u1 , f, g).
s=m+1

If m + 1 ≤ s and k ≤ ` − 1 + m, then

k − s ≤ ` − 2 ≤ ` − 1;

thus by (7.87),


κku`κ kH m−`+k−s+2 (Ω) ≤ Cs−k N (u0 , u1 , f, g).

As a consequence,


ku`κ kH m−`+k+1 (Ω) + κku`κ kH m−`+k+2 (Ω)
h m−`+k+1
X i
≤C m−s s−k + −k N (u0 , u1 , f, g) ≤ C−k N (u0 , u1 , f, g).
s=m+1

Therefore, (7.83) holds if 0 ≤ k ≤ ` + m.

3. Suppose that (7.83) holds for the case ` + (p − 1)m ≤ k ≤ ` − 1 + pm.


Then if ` + pm ≤ k ≤ ` − 1 + (p + 1)m (and k ≤ Km),

m−`+k+1
X
kaκ (0)kC s (Ω) ku`κ kH m−`+k−s+2 (Ω)
s=1
m
X m−`+k+1
X 
= + kaκ (0)kC s (Ω) ku`κ kH m−`+k−s+2 (Ω) .
s=1 s=m+1

If s ≥ m + 1, then k − s ≤ ` − 1 + pm; thus

m−`+k+1
X √ 
kaκ (0)kC s (Ω) κku`κ kH m−`+k−s+2 (Ω)
s=m+1
m−`+k+1
X
≤C m−s −(k−s) N (u0 , u1 , f, g) ≤ C−k N (u0 , u1 , f, g)
s=m+1
7.3. THE PARABOLIC REGULARIZATION OF GENERAL HYPERBOLIC EQUATIONS309

for some constant C = C kakC m (Ω×[0,T ]) , λ 0 , K). Therefore, (7.86)


with κ  1 implies that

ku`κ kH m−`+k+1 (Ω) + κku`κ kH m−`+k+2 (Ω)
√ m−`+k+1
X
≤C κ kaκ (0)kC s (Ω) ku`κ kH m−`+k−s+2 (Ω) + C−k N (u0 , u1 , f, g)
s=m+1
−k
≤ C N (u0 , u1 , f, g);

thus (7.83) holds if ` + pm < k ≤ ` + (p + 1)m. By induction, (7.83)


holds for 0 ≤ k ≤ Km if (7.84) and (7.85) hold.

Now we prove (7.84) and (7.85) for all ` ≥ 0.

1. If ` = 0, (7.84) and (7.85) trivially hold because of the regularity


assumptions on fκ and gκ . Therefore, (7.83) holds for the case ` = 0.
In particular,

ku0κ kH k (Ω) ≤ C max 1, m+1−k N (u0 , u1 , f, g).



(7.88)

2. If ` = 1, (7.84) holds trivially. For (7.85), we note that if h ∈ C 1 (∂Ω)


and v ∈ H 0.5 (∂Ω), then

khvkH 0.5 (∂Ω) ≤ CkhkC 1 (∂Ω) kvkH 0.5 (∂Ω) .

By kaκ kC k (Ω×[0,T ]) ≤ C max 1, m−k kakC m (Ω×[0,T ]) and (7.88),




∂aij ∂u0κ
κ
Ni (0)

∂t ∂xj H m+k−0.5 (∂Ω)

m+k−1
X
≤C kaκt (0)kC s+1 (∂Ω) kDu0κ kH m+k−s−0.5 (∂Ω)
s=0
m+k−1
X
≤C kaκ kC s+2 (Ω×[0,T ]) ku0κ kH m+k−s+1 (Ω)
s=0
m+k−1
X
max 1, m−s−2 max 1, s−k N (u0 , u1 , f, g).
 
≤C
s=0

for some constant C = C kakC m (Ω×[0,T ]) . If 0 ≤ s ≤ m + k − 1,

m − s − 2 ≥ −k − 1 , s − k ≥ −k − 1 and m − k − 2 ≥ −k − 1 ;
310 CHAPTER 7. Linear Hyperbolic Equations

thus by max{a, b} max{c, d} ≤ max{ac, bc, ad, bd} if a, b, c, d ≥ 0,

max 1, m−s−2 max 1, s−k ≤ max 1, m−s−2 , s−k , m−k−2 ≤−k−1 .


  


Therefore, by κ−1 ≤ 1,
ij
√ ∂aκ ∂u0κ
κ Ni (0) ≤ C−k N (u0 , u1 , f, g),

∂t ∂xj H m+k−0.5 (∂Ω)

which implies that (7.85) holds for ` = 1. This also implies that (7.83)
holds for ` = 1, and

ku1κ kH k (Ω) ≤ C max 1, m−k N (u0 , u1 , f, g).



(7.89)

3. Suppose that (7.84) and (7.85) hold for all 0 ≤ ` ≤ k − 1 for some
p ≥ 2. Then similar to (7.88) and (7.89), for all ` ≤ (p − 1) and k ≥ 0,

ku`κ kH k (Ω) ≤ C max{1, m−k−`+1 N (u0 , u1 , f, g).



(7.90)

To estimate Fpκ in H m−p+k+1 (Ω), we first note that by Liebnitz rule,

k∂tp−2 t=0 (Lκ uκ )kH m−p+k+1 (Ω)


p−2 m−p+k+1
X X h
≤C kaκ kC p+s−r−2 (Ω×[0,T ]) kurκ kH m−p−s+k+3 (Ω)
r=0 s=0
+ kbκ kC p+s−r−2 (Ω×[0,T ]) kurκ kH m−p−s+k+2 (Ω) (7.91)
i
+ kcκ kC p+s−r−2 (Ω×[0,T ]) kurκ kH m−p−s+k+1 (Ω) .

By (7.90),
p−2 m−p+k+1
X X
kaκ kC p+s−r−2 (Ω×[0,T ]) kurκ kH m−p−s+k+3 (Ω)
r=0 s=0
p−2 m−p+k+1
X X  m−p−s+r+2 p+s−k−r−2 m−k
≤ max 1,  , , N (u0 , u1 , f, g).
r=0 s=0

However, if 0 ≤ s ≤ m − p + k + 1, then

m − p − s + r + 2 > −k and p + s − k − r − 2 > −k ;

thus

max 1, m−p−s+r+2 , p+s−k−r−3 , m−k ≤ −k .



7.3. THE PARABOLIC REGULARIZATION OF GENERAL HYPERBOLIC EQUATIONS311

The other two terms in (7.91) can be estimated in the same fashion.
Therefore,

k∂tp−2 t=0 (Lκ uκ )kH m−p+k+1 (Ω) ≤ C−k N (u0 , u1 , f, g),



(7.92)

which implies that (7.84) holds for the case ` = k.

Similarly,

`−1 `−r ij
X ∂ aκ ∂urκ
Ni (0)

∂t`−r ∂xj H m−`+k+0.5 (∂Ω)

r=0
`−1 m−`+k
X X
≤C k(∂t`−r aij
κ )(0)kC s+1 (Ω) kDurκ kH m−`+k−s+0.5 (∂Ω)
r=0 s=0
`−1 m−`+k
X X
≤C kaκ kC `−r+s+1 (Ω×[0,T ]) kurκ kH m−`+k−s+2 (Ω)
r=0 s=0
`−1 m−`+k
X X
max 1, m−`+r−s−1 , `−k+s−r−1 , m−k−2 N (u0 , u1 , f, g).

≤C
r=0 s=0

If 0 ≤ s ≤ m − ` + k and 0 ≤ r ≤ ` − 1, then

m − ` + r − s − 1 ≥ −k − 1 and ` − k + s − r − 1 ≥ −k − 1 ;

thus

max 1, m−`+r−s−1 , `−k+s−r−1 , m−k−2 ≤ −k−1 .




As a consequence,

`−1 `−r ij
X ∂ aκ ∂urκ
Ni (0) ≤ C−k−1 N (u0 , u1 , f, g)

∂t`−r ∂xj H m−`+k+0.5 (∂Ω)

r=0

which implies that


κkG`κ kH m−`+k+0.5 (∂Ω) ≤ C−k N (u0 , f, g).

The estimates for the case ` = m and ` = m + 1 are similar, and are
left as an exercise.

312 CHAPTER 7. Linear Hyperbolic Equations

Estimates for δfκ and δgκ

Lemma 7.19 establishes estimates corresponding to (7.66) in the case of wave


equations. In order to show that the solution to the parabolic regularization
has κ-independent estimates, estimates corresponding to (7.70) and (7.71)
have to be established as well. Similar to (7.70) and (7.71), we have the
following

Lemma 7.20. Let δfκ and δgκ be given in Definition 7.18. Then

k∂tk (δfκ )kH m−k−1 (Ω) + k∂tk (δgκ )kH m−k−0.5 (∂Ω) ≤ Cκk−m ×
h i
× ku0 kH m+1 (Ω) + ku1 kH m (Ω) + kf kW m−1 (T ) + kgkW m−0.5 (T ) .
Ω ∂Ω
(7.93)

Proof. We recall that

m−3
X t`  
u(`+2)κ − ∂t` t=0 fκ − Lκ uκ + κ(aij κ

δfκ = κ ut ,j ),i ,
`!
`=0
m−1 `
t` ` ∂u(r+1)κ `−k ij
XX  
δgκ = κNi (∂t aκ )(0) .
`! r ∂xj
`=0 r=0

We note that for ` ≤ m − 2,

`
∂t (aij uκ , ), ≤ ∂t` t=0 (aij κ

t=0 κ t j i H m−k+1 (Ω) κ ut ,j ) H m−k+2 (Ω)

` m−k+2
X X
≤ k(∂t`−r aκ )(0)kC s (Ω) ku(r+1)κ kH m−k−s+3 (Ω)
r=0 s=0
X m−k+2
` X
max 1, m−`+r−s , k+s−r−3 , m−`+k−3 N (u0 , u1 , f, g)

≤C
r=0 s=0
≤ Ck−m−1 N (u0 , u1 , f, g). (7.94)

Therefore, by (7.80a),

m−3
X
k∂tk (δfκ )kH m−k−1 (Ω) u(`+2)κ − ∂t` (fκ − Lκ uκ m−k+1

≤C t=0 H (Ω)
`=k
m−3
X `
∂t (aij uκ , ),

+ Cκ t=0 κ t j i H m−k+1 (Ω)
;
`=k
7.3. THE PARABOLIC REGULARIZATION OF GENERAL HYPERBOLIC EQUATIONS313

thus (7.94) implies that

k∂tk (δfκ )kH m−k−1 (Ω)


m−3
X ij
(aκ (0)u(`+2)κ ,j ),i m−k−1 + Cκk−m−1 N (u0 , u1 , f, g)

≤ Cκ H (Ω)
`=k
k−m
≤ Cκ N (u0 , u1 , f, g).

Similarly,

k∂tk (δgκ )kH m−k−0.5 (∂Ω)


m−1
XX ` m−k−1
X
≤ Cκ k(∂t`−r aij
κ )(0)kC s+1 (∂Ω) kDu(r+1)κ kH m−k−s−0.5 (∂Ω)
`=k r=0 s=0
m−1
XX ` m−k−1
X
≤ Cκ kaij
κ kC `−r+s+1 (Ω×[0,T ]) ku(r+1)κ kH m−k−s+1 (Ω)
`=k r=0 s=0
m−1
XX ` m−k−1
X
max 1, m−`+r−s−1 max 1, k+s−r−1 N (u0 , u1 , f, g)
 
≤ Cκ
`=k r=0 s=0
m−1
XX ` m−k−1
X
max 1, m−`+r−s−1 , k+s−r−1 , m−`−2 N (u0 , u1 , f, g)

≤ Cκ
`=k r=0 s=0
≤ Cκk−m N (u0 , u1 , f, g),

where the last inequality follows from that m − ` + r − s − 1 ≥ k − m,


k + s − r − 1 ≥ k − m, and m − ` − 2 ≥ k − m if k ≤ ` ≤ m − 1, 0 ≤ r ≤ `,
and 0 ≤ s ≤ m − k − 1. 

1
Finally, choose γ = . Then Lemma 7.20 implies that ∂tk (δfκ ) → 0
m+1
in H m−k−1 (Ω) and ∂tk (δgκ ) → 0 in H m−k−0.5 (∂Ω). Moreover, with this
choice of γ, we have the following

Lemma 7.21. Let  = κ1/(m+1) . Then u`κ → u` in H m−`+1 (Ω) for all
0 ≤ ` ≤ m + 1.
314 CHAPTER 7. Linear Hyperbolic Equations

Proof. First of all, by (7.80a),

ku`κ − F`κ kH m−`+1 (Ω) = κ (aij



κ (0)u`κ ,j ),i H m−`+1 (Ω)

m−`+2
X
≤ Cκ kaij
κ (0)kC s (Ω) ku`κ kH m−`−s+2 (Ω)
s=0
m−`+2
X
max 1, m−s , s−2 , m−2 N (u0 , u1 , f, g)

≤ Cκ
s=0
1−m
≤ Cκ N (u0 , u1 , f, g);

thus u`κ − F`κ → 0 in H m−`+1 (Ω). The conclusion then follows from the
fact that F`κ → u` in H m−`+1 (Ω) for all 0 ≤ ` ≤ m + 1 by the property of
convolution. 
1
Remark 7.22. In Lemma 7.21 we could have chosen  = − , in which
log κ
case for all m ∈ N,

u`κ → u` in H m−`+1 (Ω),


∂tk (δfκ ) → 0 in H m−k−1 (Ω),
∂tk (δgk ) → 0 in H m−k−0.5 (∂Ω),

as κ → 0.

7.3.2 Main theorem

Theorem 7.23. Assume that ∂Ω is of class C ∞ , ∂tt + L is uniformly hy-


perbolic, aij , aij i m ij
t , b , c ∈ C (Ω × [0, T ]), and [a ] is symmetric. Then for any
u0 ∈ H m+1 (Ω), u1 ∈ H m (Ω), f ∈ FΩm−1 (T ), and g ∈ G∂Ω
m−0.5
(T ), the unique
weak solution u to the hyperbolic initial-boundary value problem (7.1a,b,c,e)
in fact belongs to W m+1 (T ; Ω) provided that u0 satisfies the compatibility
condition up to m-th order; that is,
k−1 
∂uk ∂kg X k ∂ k−` aij ∂u`

aij (0) Ni = k (0) − Ni k−`
(0) on ∂Ω
∂xj ∂t ` ∂t ∂xj
`=0

for k = 1, · · · , m. Moreover, u satisfies


kukW m+1 (T ;Ω)
h i
≤ C ku0 kH m+1 (Ω) + ku1 kH m (Ω) + kf kW m−1 (T ) + kgkW m−0.5 (T )
Ω ∂Ω
(7.77)
7.4. AN ALTERNATIVE APPROACH - THE GALERKIN METHOD315


where C = C k(a, at , b, c)kC m (Ω×[0,T ])4 , λ 0 , T .

The proof of Theorem 7.23 is essentially the same as for the case of wave
equations, which involves studying the m-times time-differentiated version
of the problem, and building up regularity using Corollary 7.10. The detail
is left as an exercise.

7.4 An alternative approach - the Galerkin method


7.4.1 Existence of weak solutions

Similar to the parabolic case, we can apply the Galerkin method to construct
a weak solution to equation (7.1a,b,c,e). In the following, we assume that
k
dk` (t)e` (t) be the finite dimensional
P
g = 0 for simplicity. Let uk (x, t) =
`=1
approximation of u(x, t) satisfying
  
uktt (t), ϕ L2 (Ω)+ B uk (t), ϕ = f (t), ϕ L2 (Ω) ∀ ϕ ∈ span(e1 , · · · , ek ) .
(7.95)

Solving the system above is equivalent to solving an ODE system


k
M`m (t)dk` (t) = fˆm (t) ,
X
dkm00 (t) + (7.96a)
`=1
dk` (0) = (u0 , e` )L2 (Ω) , dk` 0 (0) = (u1 , e` )L2 (Ω) , (7.96b)

where the matrix M`m is defined as before. The ODE system above has a
unique solution dk (t) in (0, T ) as long as M ∈ L∞ (0, T ), and it follows that
uk exists on the time interval [0, T ]. The key here is to derive a uniform
bound for uk , ukt and uktt in appropriate spaces.

The estimates for uk and ukt

Unlike the parabolic case, in order to estimate uk , we plug in ϕ = ukt instead


of uk to obtain

1d ∂uk ∂ukt
Z Z
2 ij
kukt (t)kL2 (Ω) + a (x, t) (x, t) (x, t) dx= f (x, t)ukt (x, t) dx
2 dt Ω ∂xj ∂xi Ω
∂u
Z h i
k
− bi (x, t) (x, t)ukt (x, t) + c(x, t)uk (x, t)ukt (x, t) dx . (7.97)
Ω ∂xi
316 CHAPTER 7. Linear Hyperbolic Equations

Therefore, (6.23) together with the Hölder and Young inequalities implies
that
1 dh ∂uk ∂uk
Z i
2
kukt (t)kL2 (Ω) + aij (x, t) (x, t) (x, t) dx
2 dt Ω ∂xj ∂xi
1
≤ kat (t)kL∞ (Ω) kDuk (t)k2L2 (Ω)
2
1 h i
+ kb(t)kL∞ (Ω) kDuk (t)k2L2 (Ω) + kukt (t)k2L2 (Ω) (7.98)
2
1 h i
+ kc(t)kL∞ (Ω) kuk (t)k2L2 (Ω) + kukt (t)k2L2 (Ω)
2
1h i
+ kf (t)k2L2 (Ω) + kukt (t)k2L2 (Ω) .
2
Z t
By the fundamental theorem of calculus, uk (t) = uk (0) + ukt (s) ds , so
0
hZ s i2
kuk (s)k2L2 (Ω) ≤ 2kuk (0)k2L2 (Ω) + 2 kukt (t0 )kL2 (Ω) dt0
Z0 s
≤ 2kuk (0)k2L2 (Ω) + 2s kukt (t0 )k2L2 (Ω) dt0 . (7.99)
0

Integrating (7.98) in time over the time interval (0, t), the use of (7.99) and
the uniform hyperbolicity suggest that
h
kukt (t)k2L2 (Ω) + λ 0 kDuk (t)k2L2 (Ω) ≤ C ku0 k2H 1 (Ω) + ku1 kL2 (Ω)
Z th i
2
+ kf kL2 (0,T ;L2 (Ω)) + kukt (s)k2L2 (Ω) + kDuk (s)k2L2 (Ω) ds ,
0

and we can further conclude from the Gronwall inequality that

kukt (t)k2L2 (Ω) + kDuk (t)k2L2 (Ω)


h i (7.100)
≤ C ku0 k2H 1 (Ω)+ ku1 k2L2 (Ω)+ kf k2L2 (0,T ;L2 (Ω))

for some constant C = C kakC 1 (Ω×[0,T ] , kbkL∞ (Ω×(0,T )) , kckL∞ (Ω×(0,T )) , λ 0 , T .
We note that if aij is time-independent, the dependence of C on kakC 1 (Ω×[0,T ])
can be replaced by kakL∞ (Ω) .
Using (7.99) again to obtain an upper bound for kuk (t)k2L2 (Ω) , (7.100)
then suggests that

kukt (t)k2L2 (Ω) + kuk (t)k2H 1 (Ω)


h i (7.101)
≤ C ku0 k2H 1 (Ω)+ ku1 k2L2 (Ω)+ kf k2L2 (0,T ;L2 (Ω)) .
§7.4 An alternative approach - the Galerkin method 317

The estimates for uktt

Similar to how ukt is estimated in the parabolic case, for ϕ ∈ H 1 (Ω), we


express ϕ as the sum of ϕ1 and ϕ2 such that ϕ1 ∈ span(e1 , · · · , ek ) and


(ϕ1 , ϕ2 )H 1 (Ω) = 0 to obtain uktt (t), ϕ H 1 (Ω) = (uktt (t), ϕ1 )L2 (Ω) and con-
clude that

kuktt (t)kH 1 (Ω)0 ≤ kf (t)kL2 (Ω) + C(t)kuk (t)kH 1 (Ω) , (7.102)

where C(t) = ka(t)kL∞ (Ω) + kb(t)kL∞ (Ω) + kc(t)kL∞ (Ω) . The combination of
(7.101) and (7.102) implies that

kuk kLp (0,T ;H 1 (Ω)) + kukt kLp (0,T ;L2 (Ω)) + kuktt kL2 (0,T ;H 1 (Ω)0 )
h i (7.103)
≤ C ku0 kH 1 (Ω) + ku1 kL2 (Ω) + kf kL2 (0,T ;L2 (Ω))

for some constant C = C kakC 1 (Ω×[0,T ]) , kbkC(Ω×[0,T ]) , kckC(Ω×[0,T ]) , λ 0 , T )


independent of p.

Weak convergence

In order to apply Banach-Alaoglu theorem, the underlying space has to be


reflexive. By (7.103), uktt is uniformly bounded in L2 (0, T ; H 1 (Ω)0 ), while
Duk and ukt are uniformly bounded in Lp (0, T ; L2 (Ω)) for all 1 < p < ∞.
Therefore, there exists a subsequence kj of k and u ∈ Lp (0, T ; H 1 (Ω)), such
that for all p ∈ N,

ukj * u in Lp (0, T ; H 1 (Ω)), (7.104a)


ukj t * ut in Lp (0, T ; L2 (Ω)), (7.104b)
ukj tt * utt in L2 (0, T ; H 1 (Ω)0 ), (7.104c)

as j → ∞, and u satisfies the alternative variational formulation


Z T h
i
utt (t), v(t) H 1 (Ω) + B u(t), v(t) dt
0 Z T (7.105)
∀ v ∈ L2 (0, T ; H 1 (Ω))

= f (t), v(t) L2 (Ω) dt
0

which is equivalent to (7.6) (the proof is the same as the parabolic case).
Moreover, by the property of lower semi-continuity of norms, for all p ∈ N
318 CHAPTER 7. Linear Hyperbolic Equations

(7.103) implies that the weak limit u satisfies

kukLp (0,T ;H 1 (Ω)) + kut kLp (0,T ;L2 (Ω)) + kutt kL2 (0,T ;H 1 (Ω)0 )
h i
≤ C ku0 kH 1 (Ω) + ku1 kL2 (Ω) + kf kL2 (0,T ;L2 (Ω))

for some constant C independent of p. By that lim khkLp (0,T ) = khkL∞ (0,T ) ,
p→∞
we conclude that

kukL∞ (0,T ;H 1 (Ω)) + kut kL∞ (0,T ;L2 (Ω)) + kutt kL2 (0,T ;H 1 (Ω)0 )
h i
≤ C ku0 kH 1 (Ω) + ku1 kL2 (Ω) + kf kL2 (0,T ;L2 (Ω)) ;
(7.106)
thus u ∈ L∞ (0, T ; H 1 (Ω)) with time derivatives ut ∈ L∞ (0, T ; L2 (Ω)) and
utt ∈ L2 (0, T ; H 1 (Ω)0 ).
Finally, let ζ ∈ C 2 ([0, T ]) with the property that ζ(0) = ζ 0 (0) = 1 and
ζ(T ) = ζ 0 (T ) = 0. Using v(x, t) = ζ(t)ϕ(x) for some ϕ ∈ H 1 (Ω) as a test
function in (7.105), integrating by parts twice in time we find that
Z Th
i
u(t), ζ 00 (t)ϕ L2 (Ω))+ B u(t), ζ(t)ϕ dt

− ut (0), ϕ)L2 (Ω)+ u(0), ϕ)L2 (Ω)+
0
Z Th
i
− ut (t), ζ 0 (t)ϕ L2 (Ω)) + B u(t), ζ(t)ϕ dt

= − ut (0), ϕ)L2 (Ω) +
0
Z T
∀ ϕ ∈ H 1 (Ω).

= f (t), ζ(t)ϕ L2 (Ω) dt (7.107)
0

Multiplying (7.95) by ζ, integrating by parts in time twice, and then passing


k → ∞, we obtain that
Z Th
i
u(t), ζ 00 (t)ϕ L2 (Ω))+ B u(t), ζ(t)ϕ dt

− u1 , ϕ)L2 (Ω)+ u0 , ϕ)L2 (Ω)+
0
Z Th
i
− ut (t), ζ 0 (t)ϕ L2 (Ω)) + B u(t), ζ(t)ϕ dt

= − u1 , ϕ)L2 (Ω) +
0
Z T
∀ ϕ ∈ H 1 (Ω).

= f (t), ζ(t)ϕ L2 (Ω) dt (7.108)
0

Comparing (7.107) to (7.108), we conclude that u(0) = u0 and ut (0) =


u1 . As a consequence, we establish the existence of a weak limit u to the
hyperbolic initial-boundary value problem (7.1a,b,c,e).
The detail derivation of (7.107) and (7.108) is similar to (6.19) and (6.20)
in the parabolic case, and is left as an exercise.
§7.4.2 Regularity theory revisited 319

7.4.2 Regularity theory revisited

Suppose that u0 ∈ H 2 (Ω), u1 ∈ H 1 (Ω), f ∈ L2 (0, T ; L2 (Ω)) with ft ∈


L2 (0, T ; L2 (Ω)), and aij , bi , c ∈ C 1 (Ω × [0, T ]). As for the parabolic case,
in order to show that the solution u to (7.1a,b,c,e) possesses better spatial
regularity, we tend to apply the elliptic regularity to the elliptic equation

Lu = f − utt in Ω,
∂u
aij Nj = 0 on ∂Ω ;
∂xi
thus result in the study of the regularity of utt . However, unlike the case
that one can test the parabolic equation against ut to obtain an estimate for
kut kL2 (Ω) , for the hyperbolic problem we do not have the luxury to test the
hyperbolic equations against utt in order to obtain meaningful estimates.
A naive idea is to obtain the regularity of utt by studying of the time-
differentiated version of the problem: let w = ut , and obtain the regularity
of wt by testing the time-differentiated version of the problem, a hyperbolic
equation of w, against wt . However, this process will lead to the boundary
condition
∂w ∂u
aij Ni = −aij
t Ni on ∂Ω (7.109)
∂xj ∂xj

in which the right-hand side does not make sense when a is time-dependent
and u ∈ L2 (0, T ; H 1 (Ω)). To resolve this issue, we focus on the equation
K
X
uttt − aij ut ,j ,i− aij u,j ,i = F + aij hk ,j ,i in Ω×(0, T ) , (7.110a)
  

k=1
K
∂ut ∂u X ∂hk
aij Ni + aij Ni = G− aij Ni on ∂Ω×(0, T ) ,
∂xj ∂xj ∂xj
k=1
(7.110b)
u = u0 , ut = u1 , utt = u2 on Ω×{t = 0} ,
(7.110c)

where aij are assumed to satisfy aij ξj ξi ≥ λ 0 |ξ|2 , and aij , aij , aij 1
k ∈ C (Ω ×
[0, T ]) for k = 1, · · · , K, F ∈ L2 (0, T ; L2 (Ω)), G, Gt ∈ L2 (0, T ; H −0.5 (∂Ω)),
and hk ∈ W 2 (T ; Ω) for k = 1, · · · , K. For the moment we assume that the
initial data possess the regularity u0 ∈ H 1 (Ω), u1 ∈ H 1 (Ω) and u2 ∈ L2 (Ω).
320 CHAPTER 7. Linear Hyperbolic Equations

We note that (7.110) is the first time-differentiated version of (7.1) if aij =


aij i
t , b = c = 0, F = ft , G = gt , and hk = 0 for all k, and all the higher
order time-differentiated version of (7.1) can be written in the form (7.110)
with different u, aij , aij
k and hk .

Definition 7.24. A function u ∈ L∞ (0, T ; H 1 (Ω)) with time-derivatives


ut ∈ L∞ (0, T ; H 1 (Ω)), utt ∈ L2 (0, T ; L2 (Ω)) and utt ∈ L2 (0, T ; H 1 (Ω)0 ) is
said to be a weak solution to (7.110) provided that

  
uttt (t), ϕ H 1 (Ω) + B1 ut (t), ϕ + B2 u(t), ϕ = F (t), ϕ L2 (Ω) (7.111)
K Z
aij
X
1


+ G(t), ϕ H 0.5 (∂Ω) − k hk ,j ϕ,i dx ∀ ϕ ∈ H (Ω), a.e. t ∈ (0, T ) ,
k=1 Ω

and

u = u0 , ut = u1 , utt = u2 on Ω × {t = 0} , (7.112)

where B1 ut (t), ϕ is defined by
∂ut ∂ϕ
Z
aij (x, t)

B1 ut (t), ϕ ≡ (x, t) (x) dx,
∂xj ∂xi
ZΩ
∂u ∂ϕ
aij (x, t)

B2 u(t), ϕ ≡ (x, t) (x) dx.
Ω ∂xj ∂xi
Equation (7.111) is called the variational formulation of (7.110).

Since the estimates for all K are the same, we assume that K = 1
and write h1 = h. Let {e` }∞ 2
`=1 be an orthonormal basis of L (Ω) which
k
is orthogonal in H 1 (Ω), and let uk (x, t) =
P k
dm (t)em (x) be the finite
m=1
dimensional approximation of the solution, where dkm satisfies the ODE
k h
X i
dkm000 (t) + M`m (t)dk` 0 (t) + M̃`m (t)dk` (t) = F̂m (t) + Ĝm (t) + Ĥm (t),
`=1
(7.113a)
dk` (0)=(u0 , e` )L2 (Ω) , dk` 0 (0)=(u1 , e` )L2 (Ω) , dk` 00 (0)=(u2 , e` )L2 (Ω) ,
(7.113b)

where M`m (t) = B1 (e` , em ), M̃`m (t) = B2 (e` , em ), F̂m (t) = F (t), em L2 (Ω) ,


Ĝm (t) = G(t), e` H 0.5 (∂Ω) , and
Z
Ĥm (t) = aij (t)h,j (t)ϕ,i dx.

§7.4.2 Regularity theory revisited 321

Note that (7.113a) is the same as

(ukttt , ϕ)L2 (Ω) + B1 (ukt , ϕ) + B2 (uk , ϕ) = (F, ϕ)L2 (Ω)


Z
+ hG, ϕiH 0.5 (∂Ω) − aij h,j ϕ,i dx ∀ ϕ ∈ span(e1 , · · · , ek ).

(7.114)
By the fundamental theorem of ODE, there exists a unique dk ∈ C 3 ([0, Tk ])
for some 0 < Tk ≤ T if aij , aij ∈ C(Ω × [0, T ]).
Let ϕ = uktt in (7.114), we find that

1 dh i
kuktt (t)k2L2 (Ω) + B1 ukt (t), ukt (t) = F (t), uktt (t) L2 (Ω)

2 dt
1 ∂ukt ∂ukt
Z
aij


+ t (t) (t) (t)dx + G(t), uktt (t) H 0.5 (∂Ω) (7.115)
2 Ω ∂xj ∂xi
∂h ∂uktt
Z
aij

− dx − B2 uk (t), uktt (t) .
Ω ∂xj ∂xi

The key to close the estimate is to estimate the time integral of the last two
terms on the right-hand side. Nevertheless, integrating by parts in time, we
find that
Z t 
B2 uk (s), uktt (s) ds
0
Z tZ
 s=t ∂ h ij ∂u (x, s) ∂ukt (x, s)
i
= B2 uk (s), ukt (s) − a (x, s) k dxds
s=0 0 Ω ∂s ∂xj ∂xi
h i
≤ C kDuk (t)k2L2 (Ω) + kuk (0)k2H 1 (Ω) + kukt (0)k2H 1 (Ω) + kDukt (t)k2L2 (Ω)
Z th i
+C kDuk (s)k2L2 (Ω) + kDukt (s)k2L2 (Ω) ds
0
h i Z t
≤ C ku0 kH 1 (Ω) + ku1 kH 1 (Ω) + kDukt (t)kL2 (Ω) + C kDukt (s)k2L2 (Ω) ds,
2 2 2
0

and
Z tZ
∂h ∂uktt
aij (s)
(s) (s)dxds
0 Ω ∂xj ∂xi
Z
∂h ∂ukt s=t Z t Z ∂ h ∂h i ∂u
ij kt
= a (s) (s) (s)dx − aij (s) (s) (s)dxds

Ω ∂x j ∂x i s=0 0 Ω ∂s ∂x j ∂x i
h i Z t
≤ C khkW 2 (T ;Ω)+ kukt (0)k2H 1 (Ω) + kDukt (t)k2L2 (Ω)+ C kDukt (s)k2L2 (Ω) ds.
0
322 CHAPTER 7. Linear Hyperbolic Equations

Moreover, similar to (7.20) we have


Z t
hG, uκtt iH 0.5 (∂Ω) ds
0
h i
≤ C kG(0)k2H −0.5 (∂Ω) + ku1 k2H 1 (Ω) + kGt k2L2 (0,T ;H −0.5 (∂Ω))
Z t
2 2
+ C kG(t)kH −0.5 (∂Ω) + kukt (t)kH 1 (Ω) + C kukt (s)k2H 1 (Ω) ds
0
h i
≤ C ku1 kH 1 (Ω) + kGkL2 (0,T ;H −0.5 (∂Ω)) + kGt k2L2 (0,T ;H −0.5 (∂Ω))
2 2

Z th i
2
+ kDukt (t)kL2 (Ω) + C kuktt (s)k2L2 (Ω) + kDukt (s)k2L2 (Ω) ds;
0

thus integrating (7.115) in time over the time interval (0, t), by the positive
definiteness of aij and choosing  > 0 small enough, we find that
h
kuktt (t)k2L2 (Ω) + kDukt (t)k2L2 (Ω) ≤ kuktt (0)k2L2 (Ω) + C ku0 k2H 1 (Ω) + ku1 k2H 1 (Ω)
+ kF k2L2 (0,T ;L2 (Ω)) + kGk2L2 (0,T ;H −0.5 (∂Ω)) + kGt k2L2 (0,T ;H −0.5 (∂Ω))
i Z th i
+ khkW 2 (T ;Ω) + C kuktt (s)k2L2 (Ω) + kDukt (s)k2L2 (Ω) ds.
0

Therefore, by the Gronwall inequality and that kuktt (0)kL2 (Ω) ≤ ku2 kL2 (Ω) ,
we conclude that
h
kuktt (t)kL2 (Ω)+ kukt (t)kH 1 (Ω) ≤ C ku0 kH 1 (Ω)+ ku1 kH 1 (Ω)+ ku2 kL2 (Ω)
i
+ kF kL2 (0,T ;L2 (Ω))+ kGkL2 (0,T ;H −0.5 (∂Ω))+ kGt kL2 (0,T ;H −0.5 (∂Ω))+ khkW 2 (T ;Ω)

for some constant C = C kakC 1 (Ω×[0,T ]) , kakC 1 (Ω×[0,T ]) , kakC 1 (Ω×[0,T ]) , λ 0 .
Similar to argument before, we also have
h
kukttt kL2 (0,T ;H 1 (Ω)0 ) ≤ C kukt kL2 (0,T ;H 1 (Ω)) + kuk kL2 (0,T ;H 1 (Ω))
i
+ kF kL2 (0,T ;L2 (Ω)) + kGkL2 (0,T ;H −0.5 (∂Ω)) + khkW 2 (T ;Ω)
h
≤ C ku0 kH 1 (Ω) + ku1 kH 1 (Ω) + ku2 kL2 (Ω) + kF kL2 (0,T ;L2 (Ω))
i
+ kGkL2 (0,T ;H −0.5 (∂Ω)) + kGt kL2 (0,T ;H −0.5 (∂Ω)) + khkW 2 (T ;Ω) .

Therefore, if u is a weak limit of uk in Lp (0, T ; H 1 (Ω)), then

kuttt kL2 (0,T ;H 1 (Ω)0 ) + kutt kL∞ (0,T ;L2 (Ω)) + kut kL∞ (0,T ;H 1 (Ω))
h
≤ C ku0 kH 1 (Ω) + ku1 kH 1 (Ω) + ku2 kL2 (Ω) + kF kL2 (0,T ;L2 (Ω)) (7.116)
i
+ kGkL2 (0,T ;H −0.5 (∂Ω)) + kGt kL2 (0,T ;H −0.5 (∂Ω)) + khkW 2 (T ;Ω) ,
§7.4.2 Regularity theory revisited 323

and u satisfies that for all v ∈ L2 (0, T ; H 1 (Ω)),


Z Th

 i
uttt (t), v(t) H 1 (Ω) + B1 u(t), v(t) + B2 u(t), v(t) dt
0
Z Th
∂h ∂ϕ i
Z
F (t), v(t) L2 (Ω) + G(t), v(t) H 0.5 (∂Ω) − aij (t)


= (t) dx dt
0 Ω ∂xj ∂xi

which implies that u satisfies (7.111) as well. The proof of showing u = u0


and ut = u1 at t = 0 in Section 7.4.1 can be applied to show the validity of
(7.112), so u is a weak solution to (7.110).

Lemma 7.25. Assume that aij , aij , aij 1 ij


k ∈ C (Ω × [0, T ]) satisfying a ξj ξi ≥
λ 0 |ξ|2 , u0 ∈ H 1 (Ω), u1 ∈ H 1 (Ω) and u2 ∈ L2 (Ω). Then for any given
F ∈ L2 (0, T ; L2 (Ω)), G, Gt ∈ L2 (0, T ; H −0.5 (∂Ω)), and hk ∈ W 2 (T ; Ω) for
k = 1, · · · , K, there exists a weak solution u ∈ L∞ (0, T ; H 1 (Ω)) to (7.110)
satisfying

kuttt kL2 (0,T ;H 1 (Ω)0 ) + kutt kL∞ (0,T ;L2 (Ω)) + kut kL∞ (0,T ;H 1 (Ω))
h
≤ C ku0 kH 1 (Ω) + ku1 kH 1 (Ω) + ku2 kL2 (Ω) + kF kL2 (0,T ;L2 (Ω)) (7.117)
K
X i
+ kGkL2 (0,T ;H −0.5 (∂Ω)) + kGt kL2 (0,T ;H −0.5 (∂Ω)) + khk kW 2 (T ;Ω)
k=1

for some constant C = C kakC 1 (Ω×[0,T ]) , kakC 1 (Ω×[0,T ]) , kakC 1 (Ω×[0,T ]) , λ 0 .

Now let aij = aij


t , F = ft , G = gt and hk = 0 for all k in (7.110),
0.5 (T ).
and suppose that u0 ∈ H 2 (Ω), u1 ∈ H 1 (Ω), f ∈ FΩ0 (T ) and g ∈ G∂Ω
Integrating (7.111) in time over the time interval (0, t),
∂u0 ∂ϕ
Z
aij (x, 0)
 
utt (t), ϕ L2 (Ω) − (u2 , ϕ)L2 (Ω) + B1 u(t), ϕ − dx
∂xj ∂xi
  Ω 
= f (t), ϕ L2 (Ω) − f (0), ϕ L2 (Ω) + g(t), ϕ L2 (∂Ω) − g(0), ϕ L2 (∂Ω) .

Integrating by parts in xi , the equality above implies that


   
utt (t), ϕ L2 (Ω) + B1 u(t), ϕ = f (t), ϕ L2 (Ω) + g(t), ϕ L2 (∂Ω)
   ∂u0 
+ u2 − aij (0)u0 ,j ,i −f (0), ϕ 2 + aij (0)

Ni − g(0), ϕ 2 .
L (Ω) ∂xj L (∂Ω)

Therefore, if u2 = f (0) + aij (0)u0 ,j ,i (which is (7.1a) evaluated at t = 0)



∂u0
and aij (0) Ni = g(0) (which is the first-order compatibility condition for
∂xj
324 CHAPTER 7. Linear Hyperbolic Equations

(7.1a,b,c,e)), then u is the unique weak solution to (7.1a,b,c,e). Moreover,


by (7.117)

kuttt kL2 (0,T ;H 1 (Ω)0 ) + kutt kL∞ (0,T ;L2 (Ω)) + kut kL∞ (0,T ;H 1 (Ω))
h
≤ C ku0 kH 1 (Ω) + ku1 kH 1 (Ω) + f (0) + aij (0)u0 ,j ,i L2 (Ω)

i
+ kft kL2 (0,T ;L2 (Ω)) + kgt kL2 (0,T ;H −0.5 (∂Ω)) + kgtt kL2 (0,T ;H −0.5 (∂Ω))
h
≤ C ku0 kH 2 (Ω) + ku1 kH 1 (Ω) + kft kL2 (0,T ;L2 (Ω))
i
+ kgt kL2 (0,T ;H −0.5 (∂Ω)) + kgtt kL2 (0,T ;H −0.5 (∂Ω))

for some constant C = C kakC 1 (Ω×[0,T ]) , kat kC 1 (Ω×[0,T ]) , λ 0 . The elliptic
regularity then implies that
h
kukL∞ (0,T ;H 2 (Ω)) ≤ C kf kL∞ (0,T ;L2 (Ω)) + kgkL∞ (0,T ;H 0.5 (∂Ω)) + kutt kL∞ (0,T ;L2 (Ω))
i
+ kukL∞ (0,T ;L2 (Ω))
h i
≤ C ku0 kH 2 (Ω) + ku1 kH 1 (Ω) + kf kF 0 (T ) + kgkG 0.5 (T ) .
Ω ∂Ω

If bi or c is non-zero, similar argument can be made; thus we establish the


following

Theorem 7.26. Assume that ∂Ω is of class C 2 , ∂tt + L is uniformly hy-


perbolic, aij , aij i 1 ij
t , b , c ∈ C (Ω × [0, T ]), and [a ] is symmetric. Then for any
0.5 (T ), the unique weak so-
u0 ∈ H 2 (Ω), u1 ∈ H 1 (Ω), f ∈ FΩ0 (T ), and g ∈ G∂Ω
lution u to the hyperbolic initial-boundary value problem (7.1a,b,c,e) in fact
belongs to L∞ (0, T ; H 2 (Ω)) provided that u0 satisfies
∂u0
aij (0) Ni = g(0) on ∂Ω . (7.118)
∂xj

Moreover, u satisfies

kukL∞ (0,T ;H 2 (Ω)) + kut kL∞ (0,T ;H 1 (Ω)) + kutt kL∞ (0,T ;L2 (Ω))
h i (7.119)
≤ C ku0 kH 2 (Ω) + ku1 kH 1 (Ω) + kf kF 0 (T ) + kgkG 0.5 (T )
Ω ∂Ω


for some constant C = C k(a, at , b, c)kC 1 (Ω×[0,T ])4 , λ 0 , T .

Now suppose further that u0 ∈ H 3 (Ω), u1 ∈ H 2 (Ω), f ∈ FΩ1 (T ) and


1.5 (T ), and aij ∈ C 3 (Ω × [0, T ]) (and bi = c = 0 for simplicity). By
g ∈ G∂Ω)
§7.4.2 Regularity theory revisited 325

Theorem 7.26, if u0 satisfies the first-order compatibility condition (7.118),


then u ∈ W 2 (T ; Ω) satisfies (7.119).
Let w = ut . Then we may write the second time-differentiated version
of (7.110) as
∂  ij ∂wt  ∂  ij ∂w  ∂  ij ∂u 
wttt − a − 2at = ftt + a in Ω × (0, T ) ,
∂xi ∂xj ∂xi ∂xj ∂xi tt ∂xj
∂wt ∂w ∂u
aij Ni + 2aij
t Ni = gtt − aij
tt Ni on ∂Ω × (0, T ) ,
∂xj ∂xj ∂xj
w = u1 , wt = u2 , wtt = u3 on Ω × {t = 0} ,

and the discussion above implies that the unique weak solution w satisfies

kwttt kL2 (0,T ;H 1 (Ω)0 ) + kwtt kL∞ (0,T ;L2 (Ω)) + kwt kL∞ (0,T ;H 1 (Ω))
h
≤ C ku1 kH 1 (Ω) + ku2 kH 1 (Ω) + ku3 kL2 (Ω) + kftt kL2 (0,T ;L2 (Ω))
i
+ kgtt kL2 (0,T ;H −0.5 (∂Ω)) + kgttt kL2 (0,T ;H −0.5 (∂Ω)) + kukW 2 (T ;Ω))
h i
≤ C ku1 kH 1 (Ω) + ku2 kH 1 (Ω) + ku3 kL2 (Ω) + kf kF 1 (T ) + kgkG 1.5 (T )
Ω ∂Ω


for some constant C = C kakC 2 (Ω×[0,T ]) , kat kC 2 (Ω×[0,T ]) , λ 0 . The second-
order compatibility condition
∂u1 ∂u0
Ni + aij
aij (x, 0) t (x, 0) Ni = gt (0)
∂xj ∂xj

and u3 = ft (0) + aij (0)u1 ,j ,i + aij



t (0)u0 ,j ),i then imply that u is the
unique weak solution to (7.110) (with aij = aij
t and h = 0). This observation
implies that the unique weak solution to (7.1a,b,c,e) (with bi = c = 0)
satisfies

kutttt kL2 (0,T ;H 1 (Ω)0 ) + kuttt kL∞ (0,T ;L2 (Ω)) + kutt kL∞ (0,T ;H 1 (Ω))
h i
≤ C ku0 kH 3 (Ω) + ku1 kH 2 (Ω) + kf kF 1 (T ) + kgkG 1.5 (T )
Ω ∂Ω

Since u and ut satisfies

aij u,j ,i = f − utt



in Ω × (0, T ) ,
aij u,j Ni = g, on ∂Ω × (0, T ) ,
aij ut ,j ,i = ft − uttt − aij
 
t u,j ,i in Ω × (0, T ) ,
aij ut ,j Ni = gt − aij
t u,j Ni on ∂Ω × (0, T ) ,
326 CHAPTER 7. Linear Hyperbolic Equations

by elliptic regularity,
h
kukL∞ (0,T ;H 3 (Ω)) ≤ C kf kL∞ (0,T ;H 1 (Ω))+ kutt kL∞ (0,T ;H 1 (Ω))+ kgkL∞ (0,T ;H 1.5 (∂Ω))
i
+ kukL∞ (0,T ;L2 (Ω))
h i
≤ C ku0 kH 3 (Ω) + ku1 kH 2 (Ω) + kf kF 1 (T ) + kgkG 1.5 (T )
Ω ∂Ω

and
h
kut kL∞ (0,T ;H 2 (Ω)) ≤ C kft kL∞ (0,T ;L2 (Ω)) + kuttt kL∞ (0,T ;L2 (Ω)) +
i
+ kgt kL∞ (0,T ;H 0.5 (∂Ω)) + kukL∞ (0,T ;H 2 (Ω))
h i
≤ C ku0 kH 3 (Ω) + ku1 kH 2 (Ω) + kf kF 1 (T ) + kgkG 1.5 (T ) .
Ω ∂Ω

The combination of these estimates imply that u ∈ W 3 (T ; Ω) and satisfies


h i
kukW 3 (T ;Ω) ≤ C ku0 kH 3 (Ω) + ku1 kH 2 (Ω) + kf kF 1 (T ) + kgkG 1.5 (T ) . (7.120)
Ω ∂Ω

By induction, we establish the following

Theorem 7.27. Assume that ∂Ω is of class C m+1 , ∂tt + L is uniformly


hyperbolic, aij , aij i m ij
t , b , c ∈ C (Ω × [0, T ]), and [a ] is symmetric. Then for
m−0.5
any u0 ∈ H m+1 (Ω), u1 ∈ H m (Ω), f ∈ FΩm−1 (T ), and g ∈ G∂Ω (T ),
the unique weak solution u to the hyperbolic initial-boundary value problem
(7.1a,b,c,e) in fact belongs to W m+1 (T ; Ω) provided that u0 satisfies the
compatibility condition up to m-th order; that is,
k−1 
∂uk ∂kg X k ∂ k−` aij ∂u`

ij
a (0) Ni = k (0) − Ni k−`
(0) on ∂Ω
∂xj ∂t ` ∂t ∂xj
`=0

for k = 1, · · · , m. Moreover, u satisfies


kukW m+1 (T ;Ω)
h i
≤ C ku0 kH m+1 (Ω) + ku1 kH m (Ω) + kf kW m−1 (T ) + kgkW m−0.5 (T )
Ω ∂Ω
(7.121)

for some constant C = C k(a, at , b, c)kC m (Ω×[0,T ])4 , λ 0 , T .

7.5 Hyperbolic equations with Dirichlet boundary


condition

The proof of the existence and uniqueness of the weak solution to the hy-
perbolic initial-boundary value problem (7.1a,b,c,d) is essentially the same
7.6. EXERCISES 327

as for the Neumann case, and is left to the readers. Moreover, the regularity
theory for the Dirichlet problem is much easier than the Neumann case, be-
cause of the simpleness of the compatibility conditions (7.122). Therefore,
we state without proving the following

Theorem 7.28. Assume that ∂Ω is of class C m+1 , ∂tt + L is uniformly hy-


perbolic, aij , aij i m ij
t , b , c ∈ C (Ω × [0, T ]), and [a ] is symmetric. Then for any
u0 ∈ H m+1 (Ω), u1 ∈ H m (Ω), and f ∈ FΩm−1 (T ), the unique weak solution u
to the hyperbolic initial-boundary value problem (7.1a,b,c,d) in fact belongs
to W m+1 (T ; Ω) provided that u0 satisfies the compatibility condition up to
m-th order; that is,

τ uk = 0 on ∂Ω for k = 1, · · · , m. (7.122)

Moreover, u satisfies
h i
kukW m+1 (T ;Ω) ≤ C ku0 kH m+1 (Ω) + ku1 kH m (Ω) + kf kW m−1 (T ) + kgkW m−0.5 (T )
Ω ∂Ω


for some constant C = C k(a, at , b, c)kC m (Ω×[0,T ])4 , λ 0 , T .

7.6 Exercises

Problem 7.1. Mimic the proof of the existence of weak solutions to the
parabolic initial-boundary value problem (6.6a,b,d) to show that the weak
limit u obtained in (7.104) is in fact a weak solution to (7.1a,b,c,e). Note
that you need to show that u satisfies (7.6) with u(0) = u0 and ut (0) = u1 .
Also show that (7.6) and (7.105) are equivalent.

Problem 7.2. In this problem, we study the L∞ in time H 2 in space reg-


ularity theory of hyperbolic initial-boundary value problem (7.1a,b,c,e) with
g = 0 using a different approach. Complete the following.

1. We slightly modify Problem 5.9 in Chapter 5. Let aij ∈ C 1 (Ω) be


symmetric and positive definite satisfying

aij ξj ξi ≥ λ|ξ|2
328 CHAPTER 7. Linear Hyperbolic Equations

for some constant λ > 0, and define an equivalent H 1 -norm by the


induced norm of the inner product
∂u ∂v
Z
((u, v))H 1 (Ω) ≡ (u, v)L2 (Ω) + aij dx .
Ω ∂xj ∂xi

For a given f ∈ L2 (Ω), let Kf be the unique solution to the (elliptic)


Neumann problem
∂  ij ∂u 
u− a =f in Ω,
∂xi ∂xj
∂u
aij Ni = 0 on ∂Ω .
∂xj

Show that K : L2 (Ω) → H 2 (Ω) is bounded, and K : L2 (Ω) → L2 (Ω) is


symmetric. Construct an orthonormal basis in L2 (Ω) which is orthog-
onal in the product ((·, ·))H 1 (Ω) .
k
Let {e` }∞
P
`=1 be such a basis, and uk = (u, e` )L2 (Ω) e` . Show that
`=1
uk → u in H 1 (Ω) for all u ∈ H 1 (Ω), and

kuk kH 1 (Ω) ≤ CkukH 1 (Ω) ∀ u ∈ H 1 (Ω) (7.123)

∂u
for some C > 0. Moreover, if u ∈ H 2 (Ω) satisfies aij Ni = 0 on
∂xj
∂Ω, show that uk → u in H 2 (Ω), and for some constant C > 0,
∂u
kuk kH 2 (Ω) ≤ CkukH 2 (Ω) ∀ u ∈ H 2 (Ω) with aij Ni = 0 . (7.124)
∂xj

2. Let aij (x) = aij (x, 0) in 1. Use {e` }∞ 2


`=1 as the L -orthonormal ba-
sis in the Galerkin scheme to construct the unique weak solution: let
k
dkm (t)em (x) be the finite dimensional approximation of
P
uk (x, t) =
m=1
the solution, where dk satisfies ODE (7.96) (with M and fˆ defined
using the new basis). Show that

kuktt (t)k2L2 (Ω) + λ 0 kDukt (t)k2L2 (Ω)


h i
≤ kuktt (0)k2L2 (Ω) + C ku1 k2H 1 (Ω) + ku0 k2H 1 (Ω) (7.125)
Z th i
+C kuktt (s)k2L2 (Ω) + kDukt (s)k2L2 (Ω) ds
0

for some constant C = C kakC 1 (Ω) , kat kC 1 (Ω×[0,T ]) .
§7.6 Exercise 329

3. Show that

kuktt (0)kL2 (Ω) ≤ kvkL2 (Ω) ≤ kf (0)kL2 (Ω) + Ckuk (0)kH 2 (Ω)
h i
≤ C ku0 kH 2 (Ω) + kf kF 0 (T ) (7.126)

k
using the fact that kuktt (0)k2L2 (Ω) = |dkm00 (0)|2 .
P
m=1
Hint: You can start from the following equality
k
= fˆm (0) −
X
dkm00 (0) M`m (0)dk` (0)
Z`=1h
∂uk i
= fˆm (0) − bi (x, 0) (x, 0)em (x) + c(x, 0)uk (x, 0)em (x) dx
Ω ∂xi
∂u ∂em
Z
k
− aij (x, 0) (x, 0) (x) dx .
Ω ∂xj ∂xi

4. What can you say about the case g 6= 0 using this approach?

Problem 7.3. Prove the existence and uniqueness of the solution of (7.29)
in L2 (0, T ; H m+1 (Ω)) using the fixed-point argument.

Problem 7.4. Prove Theorem 7.23.

Problem 7.5. Prove Theorem 7.28 using

1. the method of parabolic regularization;

2. the Galerkin method.

Problem 7.6. Let Ω ⊂ R3 denote an open, bounded, connected and smooth


domain. Prove existence and uniqueness of weak solutions to

utt + Dp = 0 in Ω × (0, T ] ,
div u = 0 on Ω × (0, T ] ,
pN = −∆g u on ∂Ω × [0, T ] ,
(u, ut ) = (u0 , u1 ) on Ω × {t = 0} ,

where ∆g denotes the Laplace-Beltrami operator on ∂Ω defined in (5.6), and


N denotes the outward unit-normal to ∂Ω. To do so, consider an approxi-
330 CHAPTER 7. Linear Hyperbolic Equations

mation given by the parabolic κ-problem

uκtt + Dpκ = 0 in Ω × (0, T ] ,


div uκ = 0 on Ω × (0, T ] ,
pκ N = −∆g uκ − κ∆g uκt on ∂Ω × [0, T ] ,
(u, ut ) = (u0 , u1 ) on Ω × {t = 0} .

Use a contraction mapping algorithm to establish solutions to the κ-problem,


establish κ-independent estimates, and find the limit as κ → 0. Write the
basic inequality satisfied by the weak solution.

Problem 7.7. Consider the wave equation

utt − Ddivu = 0 in T3 × (0, T ] ,


(u, ut ) = (u0 , u1 ) on T3 × {t = 0} .

1. Prove existence and uniqueness of weak solutions;

2. prove u ∈ L∞ (0, T ; H 1 (T3 )) and write the basic inequality that u sat-
isfies.

(Hint. Consider the parabolic κ-problem

utt − κD div ut − D div u = 0 in T3 × (0, T ] ,


(u, ut ) = (u0 , u1 ) on T3 × {t = 0} .)

Problem 7.8. Let Ω ⊆ R3 denote an open, bounded, connected and smooth


domain. Consider the semi-linear Sine-Gordon wave equation:

utt − ∆u + sin(u) = 0 in Ω × (0, T ] ,


u=0 on ∂Ω × (0, T ] ,
(u, ut ) = (u0 , u1 ) on ∂Ω × {t = 0} .

Use the contraction mapping theorem to find the unique weak solution to the
Sine-Gordon equation. (Hint. Consider the parabolic κ-problem

utt − κ∆ − ∆u + sin(u) = 0 in Ω × (0, T ] ,


u=0 on ∂Ω × (0, T ] ,
(u, ut ) = (u0 , u1 ) on ∂Ω × {t = 0} .)
§7.6 Exercise 331

State the compatibility conditions for the initial data which are required to
obtain a solution u ∈ L∞ (0, T ; H 2 (Ω)).
Chapter 8

More Topics on Elliptic


Equations

8.1 Elliptic equations with Sobolev class coeffi-


cients

We continue to assume that Ω ⊂ Rn is a bounded, smooth, and open subset,


and once again consider the second-order elliptic equation
n
X ∂  ij ∂u 
− a =f in Ω, (8.1a)
∂xj ∂xi
i,j=1

u=0 on ∂Ω , (8.1b)

under the assumption that aij ∈ H k (Ω) (rather than C k (Ω)) for k > n/2,
and satisfies
n
X
0 < λ0 |ξ|2 ≤ aij (x)ξi ξj ∀ ξ ∈ Rn − {0} (8.2)
i,j=1

for some positive constant λ0 . The Lax-Milgram theorem then guarantees


the existence of a weak solution u ∈ H01 (Ω), with kukH01 (Ω) ≤ Ckf kL2 (Ω) . It
remains to develop the regularity theory and establish new estimates for the
solution which depend crucially on kakH k (Ω) .
n
Theorem 8.1. For k > , f ∈ H k−1 (Ω), and aij ∈ H k (Ω) satisfying (8.2),
2
the weak solution u ∈ H01 (Ω) to (8.1) is an H k+1 (Ω)-strong solution, and
h i
kukH k+1 (Ω) ≤ C kf kH k−1 (Ω) + P(kakH k (Ω) )kf kL2 (Ω) . (8.3)

332
8.1. ELLIPTIC EQUATIONS WITH SOBOLEV CLASS COEFFICIENTS333

Proof. Let aij ij


 = η ∗ E(a ), where η is the standard mollifier defined in

Definition 2.25, and E : H k (Ω) → H k (Rn ) is the Sobolev extension operator.


For  > 0 taken sufficiently small, aij
 is strictly elliptic, and
n
X
2
λ0 |ξ| ≤ aij
 (x)ξi ξj ∀ ξ ∈ Rn .
i,j=1

Moreover, aij ij k
 → a in H (Ω) as  → 0 . By elliptic regularity, the solution

u to
n
X ∂  ij ∂u 
− a (x) (x) = (η ∗ f )(x) in Ω, (8.4a)
∂xj ∂xi
i,j=1

u = 0 on ∂Ω , (8.4b)

belongs to H ` (Ω) for all ` ∈ N . We show that u has a uniform bound in


H k+1 (Ω) and the weak limit of u in H k+1 (Ω) is indeed the solution to (8.1).
Let {χm }N
m=0 be a partition of unity subordinate to an open cover
{Um }N
m=0 of Ω, such that Um ∩ ∂Ω 6= ∅ for all m ≥ 1 , U0 ∩ ∂Ω = ∅ , and
there exist smooth H k -class diffeomorphisms ψm : Vm ≡ B(0, 1) ∩ Rn+ → Um
for all m ≥ 1 . (The maps ψm can be taken smooth whenever Ω is smooth.)
Step 1. Boundary regularity Let χ = χm , ψ = ψm , U = Um , and
V = Vm for some fixed m ≥ 1 . Multiplying the both sides of (8.4a) by χ ,
and then composing both sides with ψ , we have
n
h X ∂  ij ∂u i
− χ a ◦ ψ = (χη ∗ f ) ◦ ψ in V .
∂xj  ∂xi
i,j=1

Let yk denote the local coordinates on V , and let v  = u ◦ ψ . Then the


equation above reads
n
X ∂  ij ∂v   h i
−χ̃(y) Akj (y) ã (y)A`i (y) (y) = (χη ∗ f ) ◦ ψ (y) ,
∂yk ∂y`
i,j,k,`=1

where χ̃ = χ ◦ ψ , ãij ij
 = a ◦ ψ , and A = (Dψ)
−1 . Let J = det(Dψ) . By

Piola’s identity,
n
X ∂  ∂v  
− χ̃(y) J(y)Akj (y)A`i (y)ãij
 (y) (y)
∂yk ∂y`
i,j,k,`=1 (8.5)
h i
= J(χη ∗ f ) ◦ ψ (y) ≡ F (y) ∀y ∈ V .
334 CHAPTER 8. MORE TOPICS ON ELLIPTIC EQUATIONS

n
Let bk` = JAkj A`i ãij
P
 . Since ψ : V → U is diffeomorphism, J may be
i,j=1
assumed to have a positive lower bound, say J ≥ J0 > 0 . Then

n
X n
X
bk` ξk ξ` = Jãij k ` T 2 2
 (Ai ξk )(Aj ξ` ) ≥ J0 λ0 |A ξ| ≥ J0 λ0 c0 |ξ| , (8.6)
k,`=1 i,j,k,`=1

where c0 is the smallest eigenvalue of AAT which has a positive lower bound.
In other words, bk` is also strictly elliptic.
Letting w = χ̃v  , we rewrite (8.5) as

n n h
X ∂  ij ∂w  X ∂ χ̃ ∂v  ∂(bij χ̃,j )  i
− b =F− (bij + bji ) + v .
∂yi ∂yj ∂yi ∂yj ∂yi
i,j=1 i,j=1

Letting ∂¯α (the α-th tangential differentiation operator) act on the equation
above, we find that ∂¯α w satisfies

n h i n h i
b (∂¯α w ),j = ∂¯α F + Cαβ (∂¯β bij )(∂¯α−β w ),j
X X X
ij

,i ,i
i,j=1 i,j=1 0<β≤α
n h   i
Cαβ ∂¯β (bij + bji )χ̃,i ∂¯α−β v,j + ∂¯β (bij χ̃,j ),i ∂¯α−β v 
X X

i,j=1 0≤β≤α
(8.7)
with the boundary condition ∂¯α w = 0 on ∂V . Note that since v  ∈ H ` (Ω)
for all ` ∈ N , ∂¯α w is a classical solution to the equation above; hence a
weak solution as well. Therefore,
h
∂¯ w ≤ C ∂¯α F H 1 (V)0
α 
H 1 (V)
 
∂¯j Db∂¯|α|−` v  1 0 + ∂¯j b∂¯|α|−` v  1 0 + ∂¯j b∂¯|α|−` Dv  1 0
X
+ H (V) H (V) H (V)
0≤j≤`≤|α|
|α| |α|−` 
∂¯` Db∂¯|α|−`−j v  1 0 + ∂¯` Db∂¯|α|−`−j Dv  1 0
X X
+ H (V) H (V)
`=1 j=0
i
+ ∂¯` b∂¯|α|−`−j v  H 1 (V)0 + ∂¯` b∂¯|α|−`−j Dv  H 1 (V)0 + ∂¯` b∂¯|α|−`−j D2 v  H 1 (V)0 .

Let σ(n) = 0 if n ≥ 3 or σ(n) = 0.5 if n = 2 . Since

h i
¯ kH 1 (V)0 ≤ Cn kf gkL2 (V) + kf ∂gk
kg ∂f ¯ 2n ,
L n+2−σ(n) (V)
8.1. ELLIPTIC EQUATIONS WITH SOBOLEV CLASS COEFFICIENTS335

we find that for |α| ≤ k ,


h i
Dv ∂¯ Db 1 0 ≤ Ck∂¯k−1 DbkL2 (V) kDv  kL∞ (V) + k∂Dv ¯  k 2n
 |α|
H (V) L 2−σ (V)
h n 2k−n n+σ 2k−n−σ i
≤ Cn kbkH k (V) kv  kH
2k   2k 
k+1 (V) kv kH 1 (V) + kv kH k+1 (V) kv kH 1 (V)
2k 2k

h 2k 2k i
(Young’s inequality) ≤ Cnδ kbk k
2k−n
H (V)
+ kbk 2k−n−σ
H k (V)
kv  kH 1 (V) + δkv  kH k+1 (Ω)
≤ Cδ P(kakH k (Ω) )kf kL2 (Ω) + δkv  kH k+1 (V) ,

where P denotes a polynomial function. Similarly, for 1 ≤ ` ≤ k − 1 ,

∂¯ Db∂¯k−` Dv  1 0 ≤ Cδ P kakH k (Ω) kf kL2 (Ω) + δkv  kH k+1 (V) .


` 
H (V)

Therefore, since k∂¯α F kH 1 (V)0 ≤ CkF kH |α|−1 (V) ≤ Ckf kH |α|−1 (Ω) ,

k∂¯α w kH 1 (V) ≤ Ckf kH k−1 (Ω) + Cδ P(kakH k (Ω) )kf kL2 (Ω) + δkv  kH k+1 (V) .
X

|α|≤k

By (8.7),

− bnn ∂¯α w,nn




X h i n h i
ij ¯α  ¯ Cαβ (∂¯β bij )(∂¯α−β w ),j
X X
α
= b (∂ w ),j + ∂ F +
,i ,i
1 ≤ i, j ≤ n i,j=1 0<β≤α
(i, j) 6= (n, n)
(8.8)
n h   i
Cαβ ∂¯β (bij + bji )χ̃,i ∂¯α−β v,j + ∂¯β (bij χ̃,j ),i ∂¯α−β v  .
X X

i,j=1 0≤β≤α

The right-hand side consists of terms with (|α| + 2)-th derivatives in which
only one derivative with respect to yn is involved (the first term) or terms
with lower order mixed derivatives (all the other terms). Hence, as bnn > 0,
it follows that

k∂¯α w kH 2 (V) ≤ Ckf kH k−1 (Ω) + Cδ P(kakH k (Ω) )kf kL2 (Ω) + δkv  kH k+1 (V) .
X

|α|≤k−1

Differentiating (8.8) with respect to yn then shows that

k∂¯α w kH 3 (V) ≤ Ckf kH k−1 (Ω) + Cδ P(kakH k (Ω) )kf kL2 (Ω) + δkv  kH k+1 (V) .
X

|α|≤k−2
336 CHAPTER 8. MORE TOPICS ON ELLIPTIC EQUATIONS

A standard induction argument provides the estimate

kw kH k+1 (V) ≤ Ckf kH k−1 (Ω) + Cδ P(kakH k (Ω) )kf kL2 (Ω) + δkv  kH k+1 (V) .

which, in turn, shows that

kχm u kH k+1 (Um ) ≤ Ckf kH k−1 (Ω) + Cδ P(kakH k (Ω) )kf kL2 (Ω) + δku kH k+1 (Ω) .
(8.9)
Step 2. Interior regularity Multiplying the both sides of (8.4a) by χ0 ,
we find that in U0 ,
n n h
X ∂  ij ∂(χ0 u )  X
ji ∂χ0 ∂u
 ∂(aij
 χ0,j ) 
i
− a = χη ∗ f − (aij
 + a ) + u .
∂xj ∂xi ∂xi ∂xj ∂xi
i,j=1 i,j=1

Let Dα act on the equation above, since Dα w = 0 on ∂Ω , exactly the same


procedure as in the boundary regularity gives us

kχ0 u kH k+1 (U0 ) ≤ Ckf kH k−1 (Ω) + Cδ P(kakH k (Ω) )kf kL2 (Ω) + δku kH k+1 (U0 ) .
(8.10)
Step 3. Global regularity Combining (8.9) and (8.10), and taking δ > 0
sufficiently small, we obtain the -independent bound:
h i
ku kH k+1 (Ω) ≤ C kf kH k−1 (Ω) + P(kakH k (Ω) )kf kL2 (Ω) . (8.11)
0
Passing to the limit as 0 → 0 for a subsequence 0 we see that u → u in
H k+1−ς (Ω) for ς > 0, and that hence u is the unique solution of (8.1), and
satisfies the estimate (8.11). 

8.2 Elliptic decomposition using the divergence and


curl
8.2.1 An H 1 (Ω) elliptic estimate

Let u : Ω ⊆ R3 → R3 be a smooth vector field. Then

curlcurlu = −∆u + Ddivu or − ∆u = curlcurlu − Ddivu . (8.12)

Testing (8.12) against u and integrating by parts, we find that


X3 Z Z
2 i i 2
kDukL2 (Ω) − u,j u Nj dS = kdivukL2 (Ω) − divu(u · N)dS
i,j=1 ∂Ω ∂Ω

3
X Z
+ kcurluk2L2 (Ω) − εijk εkrs us,r ui Nj dS ,
i,j,k,r,s=1 ∂Ω
8.2. ELLIPTIC DECOMPOSITION USING THE DIVERGENCE AND CURL337

where εijk = 1(−1) if i, j, k is an even (odd) permutation of 1, 2, 3 . Since


P3
εijk εkrs = δir δks − δis δkr ,
k=1
Z 3
hX i
kDuk2L2 (Ω) = kdivuk2L2 (Ω) +kcurluk2L2 (Ω) + ui,j uj Ni −divu(u·N) dS .
∂Ω i,j=1

We first assume that u · N = 0 . Then


3
X 3 
X X2  3 X
X 2
ui,j uj Ni = Nj ∂N ui + τjα ui,α uj Ni = − ui (uj τjα )Ni,α .
i,j=1 i,j=1 α=1 i,j=1 α=1

The equality above implies


3 X
X 2 Z
kDuk2L2 (Ω) = kdivuk2L2 (Ω) + kcurluk2L2 (Ω) − ui (uj τjα )Ni,α dS
i,j=1 α=1 ∂Ω

≤ kdivuk2L2 (Ω) + kcurluk2L2 (Ω) + C kuk2L2 (Ω) + kuk2H 1 (Ω)

or
h i
kuk2H 1 (Ω) ≤ C kuk2L2 (Ω) + kdivuk2L2 (Ω) + kcurluk2L2 (Ω) , (8.13)
where C = C(|∂Ω|H 2.5 ) depends on the regularity of ∂Ω . By a density
argument, (8.13) holds for all u ∈ H 1 (Ω) with u · N = 0 on ∂Ω .

Remark 8.2. If u = 0 on ∂Ω , then the constant C is independent of the


regularity of ∂Ω . In fact, for the case u = 0 on ∂Ω , we have

kDuk2L2 (Ω) = kdivuk2L2 (Ω) + kcurluk2L2 (Ω) .

Now suppose that u · N 6= 0 . By elliptic regularity, the solution ϕ to the


problem
1 ∂ϕ
Z
∆ϕ = u · N dS in Ω , =u·N on ∂Ω
|Ω| ∂Ω ∂N
satisfies
kϕk2H 2 (Ω) ≤ Cku · Nk2H 0.5 (∂Ω) .
Let v = u − Dϕ ∈ H 1 (Ω) , then v · N = 0 and hence
h i
kvk2H 1 (Ω) ≤ C kvk2L2 (Ω) + kdivvk2L2 (Ω) + kcurlvk2L2 (Ω)

which implies
h
kuk2H 1 (Ω) ≤ C(|∂Ω|H 2.5 ) kuk2L2 (Ω) + kdivuk2L2 (Ω) + kcurluk2L2 (Ω) (8.14)
i
+ ku · Nk2H 0.5 (∂Ω) ∀ u ∈ H 1 (Ω) .
338 CHAPTER 8. MORE TOPICS ON ELLIPTIC EQUATIONS

8.2.2 An H 2 (Ω) elliptic estimate

Suppose that Ω is an H 3 class domain (with H 2.5 -class boundary ∂Ω), and
u ∈ H 2 (Ω) . Let χ` ∈ C0∞ (Rn ) denote a partition of unity subordinate
to the open cover U` , ` = 0, 1, · · · , N of Ω, such that U0 ∩ ∂Ω = ∅ and
U` ∩∂Ω 6= ∅ for all ` 6= 0 , and for each ` 6= 0 , there exist H 3 -diffeomorphisms
ϕ` : B(0, 1) ∩ Rn+ → U` ∩ Ω . Since D(χ0 u) ∈ H 1 (Ω) and ∂(χ0 u) · N = 0 on
∂Ω , (8.14) implies that
h i
kD(χ0 u)k2H 1 (Ω) ≤ C kuk2H 1 (Ω) + kdivuk2H 2 (Ω) + kcurluk2H 2 (Ω) .

Next, let χ = χ` and ψ = ψ` = ϕ−1 for some fixed ` 6= 0 . Let A =


(Dψ)−1 and v = u ◦ ψ and η = χ ◦ ψ. Then, locally, on B(0, 1) ∩ Rn+ , we see
that

3 3 h i 3
Aji (ηv)j,iα = Aji (ηv)i,j (Aji ),α (ηv i ),j ,
X X X
[div∂α (χu)] ◦ ψ = −

i,j=1 i,j=1 i,j=1
3
X 3 h
X i X3
[curl∂α (χu)]i ◦ ψ = εijk A`j (ηv)k,`α = εijk A`j (ηv)k,` − εijk (A`j ),α (ηv)k,` .

j,k,`=1 j,k,`=1 j,k,`=1

Therefore,
h i
kdiv∂α (χu)kL2 (Ω) ≤ C(|∂Ω|2.5 ) kdiv(ηu)kH 1 (Ω) + k∂AkL4 (Ω) kηvkW 1,4 (Ω)
h i
≤ C(|∂Ω|2.5 ) kdivukH 1 (Ω) + kuk 47 .
H (Ω)

Similarly,
h i
kcurl∂α (χu)kL2 (Ω) ≤ C(|∂Ω|2.5 ) kcurlukH 1 (Ω) + kuk 7 .
H 4 (Ω)

Since ∂α (χu) ∈ H 1 (Ω) , by (8.14),


h
k∂α (χu)k2H 1 (Ω) ≤ C(|∂Ω|2.5 ) k∂α (χu)k2L2 (Ω) + kdivuk2H 1 (Ω) + kcurluk2H 1 (Ω)
i
+ kuk2 7 + k∂α (χu) · Nk2H 0.5 (∂Ω) .
H 4 (Ω)

Together with (8.12), we find that


h
kuk2H 2 (Ω) ≤ C(|∂Ω|2.5 ) kuk2L2 (Ω) + kdivuk2H 1 (Ω) + kcurluk2H 1 (Ω)
i
¯ · Nk2 0.5
+ k∂u H (∂Ω) ,
8.3. THE PROBLEM OF DIVU = F IN Ω WITH U = 0 ON ∂Ω 339

¯ denotes the tangential derivative of u along ∂Ω. By induction, in


where ∂u
general we have for k ≥ 1 ,
h
kuk2H k+1 (Ω) ≤ C(|∂Ω|H ` ) kuk2L2 (Ω) + kdivuk2H k (Ω) + kcurluk2H k (Ω)
i
¯ · Nk2 k−0.5
+ k∂u ∀ u ∈ H k+1 (Ω) ,
H (∂Ω)
(8.15)
where ` = max{2.5, k + 0.5} .
If ∂Ω possesses better regularity, say ∂Ω ∈ H ` with ` = max{3.5, k+1.5} ,
then by ∂u · N = ∂(u · N) − u · ∂N ,
h
kuk2H k+1 (Ω) ≤ C(|∂Ω|H ` ) kuk2L2 (Ω) + kdivuk2H k (Ω) + kcurluk2H k (Ω)
i
+ ku · Nk2H k+0.5 (∂Ω) ∀ u ∈ H k+1 (Ω) .
(8.16)

8.3 The problem of divu = f in Ω with u = 0 on ∂Ω

In this section, we discuss a very important problem in the study of fluid


dynamics: given f ∈ L2 (Ω)/R (i.e., Ω f (x)dx = 0), find a vector field u so
R

that divu = f in Ω and u = 0 on ∂Ω .


This problem is difficult because we not only require that u · N vanishes
on ∂Ω but also that u · τα vanishes, where τ1 and τ2 span the tangent space.
If only u · N = 0 on the boundary, then the solution is very easy to find: let
p be a solution to

−∆p = f in Ω,
∂p
=0 on ∂Ω ,
∂N
and u = Dp is a solution to divu = f in Ω with u · N = 0 on ∂Ω . The
argument for obtaining a solution u ∈ H01 (Ω) is non-trivial, and we have the
following theorem.

Theorem 8.3. Let n ≥ 2 and Ω ⊂ Rn a bounded,


Z open domain with C 3 -
boundary. Then for all f ∈ L2 (Ω) so that f (x)ds = 0 , there is a u ∈

H01 (Ω) satisfying divu = f and

kukH 1 (Ω) ≤ Ckf kL2 (Ω) (8.17)

for some constant C depending on |∂Ω|W 3,∞ .


340 CHAPTER 8. MORE TOPICS ON ELLIPTIC EQUATIONS

Proof. We will provide the proof for the case n = 3. When n 6= 3, a slight
modification of this argument using differential forms is necessary.
Step 1. Existence. We look for a solution in the form

u = curlv + Dp .

Let p be denote a solution to the Neumann problem

−∆p = f in Ω, (8.18a)
∂p
=0 on ∂Ω . (8.18b)
∂N
∂p 1
We have that kpkH 2 (Ω) ≤ Ckf kL2 (Ω) and hence ∂N ∈ H 2 (∂Ω).
We must find a vector-field v ∈ H 2 (Ω) such that

curlv + Dp = 0 on ∂Ω .
1
Let g := N × Dp ∈ H 2 (∂Ω), and define B to be the unique H 2 (Ω)-
solution of
1
Z
−∆B = gdS in Ω,
|Ω| ∂Ω
∂B
=g on ∂Ω .
∂N
Let d(x) = inf |x − y| denote the distance function to the boundary. Since
y∈∂Ω
∂Ω is C 3 , d is three times differentiable near the boundary and −Dd = N
on the boundary. Let ξ : [0, ∞) → [0, ∞) be a smooth function with (small)
compact support so that ξ(0) = 1 and ξ 0 (0) = 0 . We claim that
 
v(x) = B(x) − ξ d(x) B x − d(x)Dd(x)

satisfies curlv = −Dp on ∂Ω .


To see this, we differentiate and obtain

v,jk (x) = B,jk (x) − ξ 0 d(x) d,j (x)B k x − d(x)Dd(x)


 

3
X h i
ξ d(x) B,`k x − d(x)Dd(x) δj` − d,j (x)d,` (x) − d(x)d,j` (x) .


`=1

For x ∈ ∂Ω , d(x) = 0 . Since ξ 0 (0) = 0 , we find that


3
X ∂B k
v,jk (x) = B,`k (x)d,j (x)d,` (x) = Nj = g k Nj ∀ x ∈ ∂Ω .
∂N
`=1
8.3. THE PROBLEM OF DIVU = F IN Ω WITH U = 0 ON ∂Ω 341

Therefore, on the boundary ∂Ω ,


3
X 3
X 3
X
i
(curlv) = εijk v,jk = k
εijk g Nj = εijk εkrs p,s Nr Nj
j,k=1 j,k=1 j,k,r,s=1
3
X ∂p
= (δir δjs − δis δjr )p,s Nr Nj = Ni − p,i N · N = −p,i ,
∂N
j,r,s=1

∂p
where the condition = 0 is used to conclude the last equality.
∂N
Step 2. Estimate. First we note that by elliptic regularity B satisfies

kBk2H 2 (Ω) ≤ C(|∂Ω|W 2,∞ )|g|2H 0.5 (Γ) ≤ C(|∂Ω|W 2,∞ )kpk2H 2 (Ω)
≤ C(|∂Ω|W 2,∞ )kf k2H 1 (Ω)

and hence v satisfies

kvk2H 2 (Ω) ≤ C(|∂Ω|W 3,∞ )kBk2H 2 (Ω) ≤ C(|∂Ω|W 3,∞ )kf k2H 1 (Ω) ,

where the dependence of |∂Ω|W 3,∞ comes from the estimate of D3 d . This
estimate, together with the estimate for p, leads to (8.17). 

q
Remark Z 8.4. In general, if ∂Ω is locally Lipschitz, then for all f ∈ L (Ω)
so that f (x)dx = 0 , there exists u ∈ W01,q (Ω) satisfying divu = f and

kuk2W 1,q (Ω) ≤ Ckf k2Lq (Ω) ,

where C depends on Ω .

Corollary 8.5. Let Ω be a domain 3 Then for all f ∈


Z with C -boundary.
Z
2 0.5
L (Ω) and g ∈ H (Γ) so that f (x)ds = g(x) · N(x)dS , there is a
Ω Γ
u ∈ H 1 (Ω) so that divu = f in Ω and u = g on Γ with the estimate
h i
kuk2H 1 (Ω) ≤ C kf k2L2 (Ω) + kgk2H 0.5 (∂Ω)

for some constant C depending on |∂Ω|W 3,∞ .

8.3.1 Inequalities associated with the normal and tangential


decomposition of vector fields on ∂Ω
The regularity of ∂Ω

Definition 8.6. A (bounded) domain Ω ⊆ Rn is said to be of class C k if


∂Ω is an n − 1 dimensional C k sub-manifold of Rn , or equivalently,
342 CHAPTER 8. MORE TOPICS ON ELLIPTIC EQUATIONS

N
(1) there are open sets O` ⊂ Rn such that ∂Ω ⊆
S
O` ;
`=1

(2) there are maps ϕ` : U` := ∂Ω ∩ O` → Rn−1 so that ψ` := ϕ−1


` :
ϕ` (U` ) → Rn is a C k injective immersion, that is, for each `, ψ` is
C k , one-to-one, and for a given coordinate (y 1 , · · · , y n−1 ) in ϕ` (U` ),
the set of vectors {ψ`,1 (y), · · · , ψ`,n−1 (y)} are linearly independent for
all y ∈ ϕ` (U` );

(3) and each transition map ϕ`1 ◦ ϕ−1


`2 : U`1 ∩ U`2 → R
n−1 is C r for some

r ≥ k.

A domain Ω is smooth if Ω is a C k -domain for all k > 0.

Given a local chart (U, ϕ) on ∂Ω at a point y ∈ ∂Ω with local coordinates


yα, α = 1, ..., n − 1, let Ty ∂Ω denote the tangent space to ∂Ω at the point y:

Ty ∂Ω = {v ∈ Rn |v · N (y) = 0} .

Each such tangent space is diffeomorphic to Rn−1 with n − 1 linearly inde-


pendent tangent vectors spanning the space. The natural basis of tangent
vectors is given by the partial derivatives of ψ:
 ∂ 
:= ψ,α (y) α = 1, ..., n − 1 ,
∂y α y

which condition (2) in Definition 8.6, are linearly independent.


For a function f defined on ∂Ω, the partial derivative of f with respect
to y α at y, denoted by f,α (y), is defined as

∂f ∂(f ◦ ψ)
f,α (y) = α
(y) := . (8.19)
∂y ∂y α

Throughout, Greek indices run from 1 to n − 1, while Latin indices run from
1 to n.
Having established a local coordinate system y α , α = 1, ..., n − 1 on
the boundary ∂Ω, we extend the coordinate system to an n-dimensional
neighborhood of ∂Ω. For y n ∈ R, define the map Ψ : ϕ(U) × R → Rn by

Ψ(y 1 , · · · , y n−1 , y n ) = ψ(y 1 , · · · , y n−1 ) + y n N (ψ(y 1 , · · · , y n−1 )) , (8.20)


8.3. THE PROBLEM OF DIVU = F IN Ω WITH U = 0 ON ∂Ω 343

where N denotes the outward pointing unit normal on ∂Ω. With the diame-
ter of O taken sufficiently small, Ψ : V = Φ(O) → O is bijection mapping the
curvilinear cooridantes (y1 , ..., yn ) onto the standard Cartesian coordinates
(x1 , ..., xn ).

Ψ Ω
Φ(O+ ) ∂Ω
ψ = ϕ−1
O+
(y1 , · · · , yn−1 ) ∈ "n−1
ϕ
y ∈ ∂Ω
Φ = Ψ−1

Let p ∈ O+ = O ∩Ω. By assumption there exists a unique q = q(p) ∈ ∂Ω


so that dist(p, q) = dist(p, ∂Ω), and

*
pq = dist(p, q)N (q) .

Suppose p = Ψ(y 1 , · · · , y n ), then q = Ψ(y 1 , · · · , y n−1 , 0). Given y ∈ O+ , we


define the tangent vectors
 ∂ 
:= Ψ,i (y) .
∂y i y

Next, we define
n := N ◦ ψ .

It follows that
 ∂  h i
n
:= ψ,α + y n ,α (y 1 , · · · , y n−1 ) ; (8.21a)
∂y α y
 ∂ 
:= n(y 1 , · · · , y n−1 ) . (8.21b)
∂y n y

N
S
By adding an open set O0 ⊂⊂ Ω, we may assume that Ω ⊆ O` . Let
`=0
{χ` }N
`=0 be a partition of unity (of Ω) subordinate to O` . We will further
assume that spt(χ` ) ⊂⊂ O` , O` is smooth for all `, and ∂O` is transverse to
{yn = 0} (and equivalently, ∂Φ(O` ) is transverse to ∂Ω).
Let f be a function defined on Ω, and f` = χ` f . By the properties of
N N
{χ` }N
P P
`=0 , f = (χ` f ) = f` . For 1 ≤ i ≤ n, f` ◦ Ψ` is defined on Φ(O` ),
`=0 `=0
344 CHAPTER 8. MORE TOPICS ON ELLIPTIC EQUATIONS

or in other words, f` ◦ Ψ` is a function of y. Let A` = (Dy Ψ` )−1 , by the


chain rule,
∂f` ∂(f` ◦ Ψ` )
◦ Ψ` = (A` )ji = (A` )αi f`,α + (A` )ni f`,n . (8.22)
∂xi ∂yj
Important notation: For a given function f , we use f,i to denote the
derivative of f with respect to the Cartesian coordinates xk , and we use
f,α and f,n to denote the derivative of f with respect to the curvilinear
coordinates yk .

Tangential and normal derivatives

Decomposing vectors on ∂Ω in 2-D. Suppose that n = 2. We begin with the


definition of two important geometric quantities: in each chart (U, ϕ), let
g = ψ,1 ·ψ,1 and b = n,1 ·ψ,1 denote the first and second fundamental forms,
respectively. Then
ψ 2 ,1 ψ 1 ,1
n = (− √ , √ ) , and n,1 = g −1 bψ,1 .
g g
By the definition of Ψ,
2 
ψ,1 2 
ψ,1
y2 b
 
1 1 1
 ψ,1 + y2 n,1 − √g   ψ,1 (1 + g
) −√ 
g 
Dy Ψ =  = ;
  
1
ψ,1   2 1
ψ,1

2 2 y2 b 
ψ,1 + y2 n,1 √ ψ,1 (1 + ) √
g g g
g + y2 b
hence, det(Dy Ψ) = √ and
g
 1
ψ,1 2
ψ,1

√ √ √
g  g g 
A=  .
g + y2 b 
2 y2 b 1 y2 b 
−ψ,1 (1 + ) ψ,1 (1 + )
g g
Consequently,
1
τi1 := A1i = ψ,1 // ∂Ω , and A2i = ni .
g + y2 b
and (8.22) implies that the gradient can be decomposed into tangential and
normal derivatives:
∂f
◦ Ψ = τi1 f,1 + ni f,n ,
∂xi
8.3. THE PROBLEM OF DIVU = F IN Ω WITH U = 0 ON ∂Ω 345

where f,1 = ∂f /∂y1 and f,n = ∂f /∂y2 .


Decomposing vectors on ∂Ω in n-D. Let eα = (0, · · · , 0, 1, 0, · · · , 0)T
be the α-th tangent vector on {yn = 0}, and en = (0, · · · , 0, −1)T is the
outward pointing normal to {yn ≥ 0}. Define τ α = AT eα and ν = AT en
(where A = (Dy Ψ)−1 ). By setting Gαβ = Ψ,α ·Ψ,β and G αβ = [Gαβ ]−1 , we
find that

τ α = AT eα = G αβ Ψ,β and ν = AT e n = n . (8.23)

Note that τ α and ν are also defined away from y n = 0, and τα · ν = 0 for all
α = 1, · · · , n − 1.
By (8.22), for a given function f defined in O,

∂f ∂(f ◦ Ψ) ∂(f ◦ Ψ) ∂(f ◦ Ψ)


◦ Ψ = Aji = Aαi + Ani .
∂xi ∂y j ∂y α ∂y n
Since Aαi = τiα and Ani = ni , in Φ(O) we find that

∂(f ◦ Ψ) ∂(f ◦ Ψ)
(Dx f ) ◦ Ψ = τ α +n = τ α f,α + nf,n . (8.24)
∂y α ∂y n

Some useful inequalities

The mean curvature H is defined as the trace of the second fundamental


form bαβ so that in a local chart (U, ψ),

g αβ bαβ
H ◦ψ = , bαβ = −ψ,αβ ·n .
n−1
Lemma 8.7. Suppose that (O, Ψ) is a local chart in a neighborhood of y ∈
∂Ω with coordinates (y α , y n ). If u ∈ H s (Ω) ∩ H01 (Ω) for s > 2.5, then on
∂Ω (or on {yn } = 0 in Ψ−1 (O)),

(u,ij Ni − u,ii Nj ) ◦ ψ = τjα u,αn − (n − 1)(H ◦ ψ)nj u,n ,

where H is the mean curvature of ∂Ω.

Proof. By (8.23), for y n ≥ 0, τ α = G αβ Ψ,β . Therefore, by (8.24),

u,ij ◦ Ψ = nj (ni u,nn + τiα u,αn + ni,n u,n + τi,n


α
u,α )
+ τjβ (τiα u,αβ + ni u,βn + τi,β
α
u,α + ni,β u,n ) .
346 CHAPTER 8. MORE TOPICS ON ELLIPTIC EQUATIONS

Using this identity together with τ α · n = 0 and u,α = 0 if yn = 0, we find


that
h i
u,ij Ni − u,ii Nj ◦ ψ = τjβ (ni u,βn + τi,β
α
u,α + ni,β u,n )ni − τiβ (τiα u,αβ + τi,β
α
u,α + ni,β u,n )nj

= τjα u,αn − nj τiβ ni,β u,n .

The result then follows from the fact that

τiβ ni,β |yn =0 = g αβ ψ,α


i
ni,β = g αβ bαβ = (n − 1)H ◦ ψ .


Corollary 8.8. Given u ∈ H s (Ω) ∩ H01 (Ω) for s > 2.5,


Z Z ∂u 2
[u,ij Ni − u,ii Nj ]u,j dS = (n − 1) H dS .

∂Ω ∂Ω ∂N

Proof. By the definition of the surface integral,


Z N Z
X
[u,ij Ni − u,ii Nj ]u,j dS = χ` [u,ij Ni − u,ii Nj ]u,j dS
∂Ω `=1 ∂Ω
N Z
X h i p
= χ` [u,ij Ni − u,ii Nj ]u,j ◦ ψ` det(g` ) dy1 · · · dyn−1 .
`=1 ϕ(U` )

By Lemma 8.7,
h i
u,ij Ni − u,ii Nj ◦ ψ` = τjα u,αn − (n − 1)(H ◦ ψ)nj u`,n .

Therefore, by u`,i ◦ ψ` = ni u`,n on {yn = 0},

Z N Z
X p
[u,ij Ni − u,ii Nj ]u,j dS = −(n − 1) [(χ` H) ◦ ψ` ]u2`,n det(g` )dy1 · · · dyn−1 .
∂Ω `=1 ϕ(U` )

The corollary is immediately proved from this equality. 

Corollary 8.9. If u ∈ H 2 (Ω) ∩ H01 (Ω), then


Z ∂u 2
2
kD uk2L2 (Ω) = k∆uk2L2 (Ω) + (n − 1) H dS . (8.25)

∂Ω ∂N

Proof. We first establish the identity is valid for all u ∈ C ∞ (Ω) ∩ H01 (Ω); a
density argument then completes the proof.
8.3. THE PROBLEM OF DIVU = F IN Ω WITH U = 0 ON ∂Ω 347

Let u ∈ C ∞ (Ω) ∩ H01 (Ω). Integrating by parts, we find that


Z Z
2 2
k∆ukL2 (Ω) = |∆u(x)| dx = u,ii (x)u,jj (x)dx
ZΩ ZΩ h i
= |D2 u(x)|2 dx − u,ij (x)Ni − u,ii (x)Nj u,j (x)dS
Ω ∂Ω

and the conclusion follows from Corollary 8.8. 

Corollary 8.10. There is a constant C = C(Ω) so that


h i
kuk2H 2 (Ω) ≤ C kDuk2L2 (Ω) + k∆uk2L2 (Ω) ∀ u ∈ H 2 (Ω) ∩ H01 (Ω) . (8.26)

∂u 2
Z ∂u
Proof. We only need to estimate H dS . Since = |Du · N | ≤ |Du|,

∂Ω ∂N ∂N
Z ∂u 2
H dS ≤ C(Ω)kDuk2L2 (∂Ω) ≤ C(Ω)kDuk2H 0.25 (∂Ω) ≤ kDuk2H 0.75 (Ω) .

∂Ω ∂N

By (2.48) and Young’s inequality, we conclude that


Z ∂u 2
H dS ≤ Cδ kDuk2L2 (Ω) + δkDuk2H 1 (Ω) .

∂Ω ∂N

The identity (8.26) then follows from (8.25) and taking δ > 0 small enough.


Corollary 8.11. The norm kuk := kDukL2 (Ω) + k∆ukL2 (Ω) is an equivalent
norm in H 2 (Ω) ∩ H01 (Ω).

Remark 8.12. Similar to the proofs of Lemma 8.7 and Corollary 8.10, for
every multi-index α,

Z h i |α|+1
X
α α α
(D u),jj Ni − (D u),ij Nj (D u),i dS ≤ C kD|β| uk2L2 (∂Ω) .


∂Ω |β|=1

for all u ∈ H |α|+s (Ω) ∩ H01 (Ω). Therefore, one important conclusion of this
inequality is that the norm kuk := kukH k+1 (Ω) + k∆ukH k (Ω) is an equivalent
norm in H k+2 (Ω) ∩ H01 (Ω), or more precisely, there are constants C1 and
C2 so that for all u ∈ H k+2 (Ω) ∩ H01 (Ω),

C1 kuk2H k+2 (Ω) ≤ kuk2H k+1 (Ω) + k∆uk2H k (Ω) ≤ C2 kuk2H k+2 (Ω) . (8.27)
348 CHAPTER 8. MORE TOPICS ON ELLIPTIC EQUATIONS

8.4 Vector fields with given vorticity, divergence,


and normal trace

In this section, we are concerned with obtaining a vector field u whose


vorticity curlu , divergence divu , and the normal trace u · N are given. In
other words, we look for u which satisfies

curlu = f in Ω, (8.28a)
divu = g in Ω, (8.28b)
u·N=h on ∂Ω . (8.28c)

Z
For solvability reason, we need to assume that divf = 0 and gdx =
Z Ω
hdS .
∂Ω
Let ϕ solve the elliptic equation ∆ϕ = g in Ω with Neumann boundary
∂ϕ
condition = h on ∂Ω . Then v = u − Dϕ solves
∂N

curlv = f in Ω, (8.29a)
divv = 0 in Ω, (8.29b)
v·N=0 on Ω. (8.29c)

As long as we can solve (8.29), a solution to (8.28) can be found by adding


v and Dϕ .

8.4.1 The case Ω = R3+

We first start the discussion for the case Ω = R3+ with N = (0, 0, −1) .

Existence:

By taking the curl of (8.29a), we find that −∆v = curlf in R3+ . Since
v3 = −v · N = 0 , v3 solves a Poisson equation with Dirichlet boundary
condition; hence v3 can be solved. Once v3 is obtained, by (8.29a) again, we
8.4. VECTOR FIELDS WITH GIVEN VORTICITY, DIVERGENCE, AND NORMAL TRACE349

find that
∂v1 ∂v1 ∂v3 ∂v3
=− =− − (curlv)2 = − − f2 on R2 × {x3 = 0} ,
∂N ∂x3 ∂x1 ∂x1
(8.30a)
∂v2 ∂v2 ∂v3 ∂v3
=− =− + (curlv)1 = − + f1 on R2 × {x3 = 0} .
∂N ∂x3 ∂x2 ∂x2
(8.30b)
In other words, v1 and v2 satisfy some Poisson equations with Neumann
boundary conditions (8.30); thus we can obtain v1 and v2 via solving these
Neumann problems. Note that v1 and v2 are unique up to a constant.
Even though we can solve v = (v1 , v2 , v3 ) via Dirichlet or Neumann
problems stated above, it is not clear that if v really satisfies (8.29a-b) .
However, by tangentially differentiating the Neumann boundary condition
(8.30), that is, differentiating with respect to x1 or x2 ,
∂ h ∂v1 ∂v2 i ∂ ∂v1 ∂ ∂v2
+ = +
∂x3 ∂x1 ∂x2 ∂x1 ∂x3 ∂x2 ∂x3
∂ 2 v3 ∂ 2 v3 ∂ 2 v3
= 2 + 2 + (curlf )3 = − 2 on R2 × {x3 = 0}
∂x1 ∂x2 ∂x3
which is the same as saying that
∂divv
=0 on R2 × {x3 = 0} . (8.31)
∂N
Moreover, since −∆v = curlf , ∆(divv) = 0 . In other words, divv is Har-
monic satisfying Neumann boundary condition (8.31). Therefore, divv can
only be a constant. If we insist that v ∈ H 1 (R3+ ) , then the constant must
be zero. Therefore, (8.29b) is valid.
Since divv = 0 , curlcurlv = curlf or curl(curlv − f ) = 0 in R3+ . Then
there exists a (harmonic) scalar potential ϕ such that curlv − f = Dϕ in
∂ϕ
R3+ . By (8.30b), 0 = (curlv − f )1 = on R2 × {x3 = 0} and similarly,
∂x1
∂ϕ
= 0 on R2 × {x3 = 0} . Therefore, ϕ is a constant on the boundary.
∂x2
By assuming that f ∈ L2 (R3+ ) and v ∈ H 1 (R3+ ) , ϕ is constant in R3+ ; hence
(8.29a) is valid.

Uniqueness

Suppose that both u, v ∈ H 1 (R3+ ) satisfy (8.29). Then w = u − v is ir-


rotational, divergence-free, and has zero normal trace. As a consequence,
350 CHAPTER 8. MORE TOPICS ON ELLIPTIC EQUATIONS

∂ϕ
there exists a harmonic scalar potential ϕ such that w = Dϕ and = 0.
∂N
Therefore, ϕ = c for some constant c , hence u − v = 0 .

Estimates

Suppose that f ∈ H k (R3+ ) for some k ∈ N . We note that (curlf )3 ∈


H k−1 (R3+ ) . By elliptic regularity, kv3 kH k+1 (R3+ ) ≤ Ck(curlf )3 kH k−1 (R3+ ) ≤
Ckf kH k (R3+ ) ; thus

h ∂v3 i
kDv1 kH k (R3+ ) ≤ C k(curlf )1 kH k−1 (R3+ ) + f2 + ≤ Ckf kH k (R3+ ) ,

∂x1 H
k−0.5 2
(R ×{0})
h ∂v3
i
kDv2 kH k (R3+ ) ≤ C k(curlf )2 kH k−1 (R3+ ) + f1 − ≤ Ckf kH k (R3+ ) .

∂x2 H k−0.5 (R2 ×{0})

8.4.2 The case Ω is a bounded smooth domain

Now suppose that Ω is a bounded smooth domain of R3 . We first recall


from Section 8.3.1 the maps Ψ and ψ , and let Gαβ ≡ Ψ,α · Ψ,β denote the
metric tensor induced by Ψ , with G αβ denoting the inverse of Gαβ .
In a local coordinate system near the boundary, ∂j = Nj ∂N + G αβ Ψj,α ∂β ;
hence on ∂Ω ,

3
X 3
X h i
k i
curlw · N = εijk w,j N = εijk Nj w,n
k
+ g αβ ψ,α
j k
w,β Ni
i,j,k=1 i,j,k=1
3
X 2
X 2
X
j
= εijk ψ,α Ni g αβ w,β
k
= (N × ψ,α ) · g αβ w,β
i,j,k=1 α,β=1 α,β=1
2
X √
= g(δ − α)g αβ g γδ ψ,γ
k k
w,β ,
α,β,γ,δ=1

2 √
g(δ − α)g γδ ψ,γ to establish the last equality.
P
where we use N × ψ,α =
γ,δ=1
Therefore, if Ptan denotes the projection onto the tangent plane of ∂Ω , then
on ∂Ω ,


g(g 11 g 22 − g 12 g 12 )(ψ,2 · w,1 − ψ,1 · w,2 ) = curlw · N

⇔ ψ,2 · (Ptan w + wn N),1 − ψ,1 · (Ptan w + wn N),2 = gcurlw · N

⇔ wα,1 g2α + wα Γβ1α gβ2 − wα,2 g1α − wα Γβ2α gβ1 = gcurlw · N ;
8.4. VECTOR FIELDS WITH GIVEN VORTICITY, DIVERGENCE, AND NORMAL TRACE351

thus if DX Y denotes the covariant derivative of Y in the direction X , then


ψ,2 · Dψ,1 (Ptan w) − ψ,1 · Dψ,2 (Ptan w) = gcurlw · N , on ∂Ω . (8.32)

Similarly, we also have that

3 h
X  2
X  2
X i
divw = Nk wn Nk + wα Ψk,α + G αβ Ψk,α (wn Nk + wγ Ψk,γ ),β
,n
k=1 α=1 α,β,γ=1
3
X 2
X
= wn,n + G αβ Ψk,α (wγ,β Ψk,γ + wγ Ψk,βγ + Nk,β wn + Nk wn,β ) ;
k=1 α,β=1

thus if div∂Ω denotes the divergence operator on ∂Ω , then

2 h 2 i
Γβαβ wα + 2Hwn + wn,n
X X
divw = wα,α +
α=1 β=1

= div∂Ω (Ptan w) + 2Hwn + wn,n on ∂Ω . (8.33)

The case that Ω is convex

If Ω is a convex domain (which implies that H ≥ 0), we consider the following


elliptic problem

−∆w = f in Ω, (8.34a)
Ptan w = 0 on ∂Ω , (8.34b)
wn,n + 2Hwn = 0 on ∂Ω , (8.34c)

∂wn
here we recall again that wn,n = . We will look for solutions w ∈
∂N
Hτ1 (Ω) , where
n o
Hτ1 (Ω) ≡ w ∈ H 1 (Ω) Ptan w = 0 .

We note that Hτ1 (Ω) is a Banach space.


In order to solve (8.34), we try to find a weak formulation first, and this
amounts to compute

3 Z
X
i i j
− w,j ϕ N dS ∀ ϕ ∈ Hτ1 (Ω) .
i,j=1 ∂Ω
352 CHAPTER 8. MORE TOPICS ON ELLIPTIC EQUATIONS

However, on ∂Ω (that is, yn = 0),


3
X 3 h
X  2
X 
i
w,j Nj ϕi = j i
N wn N + wα Ψi,α
,n
i,j=1 i,j=1 α=1
2
X i
αβ
+ g Ψj,α (wn Ni + i
wσ ψ,σ ),β j i
N (ϕn N )

yn =0
α,β,σ=1

= wn,n ϕn on ∂Ω ,

we have the following

Definition 8.13. A vector-valued function w ∈ Hτ1 (Ω) is said to be a weak


solution to (8.34) if
Z Z Z
Dw · Dϕdx + 2 Hwn ϕn dS = f · ϕdx ∀ ϕ ∈ Hτ1 (Ω) . (8.35)
Ω ∂Ω Ω

Since H ≥ 0 , by (??) we know that the left-hand side of (8.35) obviously


defines a bounded, coercive bilinear form on Hτ1 (Ω) × Hτ1 (Ω) ; hence by the
Lax-Milgram theorem, there exists a unique w ∈ Hτ1 (Ω) satisfying (8.35).
If f ∈ H k (Ω) and Ω is of class H k+2 , similar techniques used to prove the
elliptic regularity imply that

kwkH k+2 (Ω) ≤ Ckf kH k (Ω) .

In this case, w solves (8.34) in the strong sense.


Now suppose that f is divergence-free. Then −∆divw = 0 in Ω . More-
over, (8.33) and (8.34b-c) imply that divw = 0 on ∂Ω . As a consequence,
divw = 0 in Ω , and this further implies that curlcurlw = f . In addi-
tion, (8.32) and (8.34b) imply that curlw · N = 0 on ∂Ω ; hence the vector
v = curlw solves (8.29) and

kvkH k+1 (Ω) ≤ Ckf kH k (Ω) .

The case that Ω is not convex

If Ω is not convex, H might be negative on some portion of ∂Ω , so there


might be no solution to (8.34) (see Exercise Problem 5.1 in Chapter 5 for
an ill-posed Robin problems). However, from the analysis above, we only
need to find a divergence-free vector w so that −∆w = f in Ω . It does
8.5. COMPENSATED COMPACTNESS AND THE DIV-CURL LEMMA353

not matter whether (8.34b) and (8.34c) hold or not. Therefore, for a given
divergence-free vector f ∈ L2 (Ω) , we can extend it so that the extension f˜
is still divergence-free and is defined on the convex domain, say B(0, R) for
some R > 0 large enough. We then consider the problem on B(0, R) instead
of Ω , or to be more precise, we consider (8.34) with B(0, R) replacing Ω .
Then the unique solution w ∈ H 2 (B(0, R)) provides us a way of constructing
a solution to (8.29), and we have

kvkH 1 (Ω) ≤ kcurlwkH 1 (Ω) ≤ Ckf kL2 (Ω) .

If f ∈ H k (Ω) with divf = 0 , the divergence-free extension f˜ in general


might not belong to H k (Ω) as well, so we cannot apply the procedure of
obtaining the elliptic regularity to show that v ∈ H k+1 (Ω) . However, we
can use the divergence and curl estimate (8.16) to obtain that
h i
kvkH k+1 (Ω) ≤ C kvkL2 (Ω) + kcurlvkH k (Ω)
h i
≤ C kwkH 1 (Ω) + k∆wkH k (Ω) ≤ Ckf kH k (Ω) .

Thus taking the scalar potential ϕ into account, we prove

Theorem 8.14. Suppose that Ω ⊆ R3 is a bounded smooth domain. Then


there exists a unique solution u ∈ H k+1 (Ω) to (8.28) if f ∈ H k (Ω) is
divergence-free,
Z g ∈Z H k (Ω) and h ∈ H k+0.5 (∂Ω) satisfies the solvability
condition gdx = hdS . Moreover, u satisfies the following estimate
Ω ∂Ω
h i
kukH k+1 (Ω) ≤ C kf kH k (Ω) + kgkH k (Ω) + khkH k+0.5 (∂Ω) .

8.5 Compensated compactness and the div-curl lemma

Lemma 8.15. Suppose vk * v and wk * w both in L2 (Ω) , and divvk and


curlwk are compact in H −1 (Ω) . Then vk · wk → v · w in D0 (Ω) .

Before proving Lemma 8.15, let us examine why this should be true.
Suppose divvk and curlwk both vanish. Then vk = curluk (v = curlu) for
some H 1 -vector field uk (u) and wk = Dpk (w = Dp) for some H 1 -scalar pk
(p). Therefore, for ϕ ∈ D(Ω) ,
Z Z
vk · wk ϕdx = curluk · Dpk ϕdx = −hcurluk , pk Dϕi ,
Ω Ω
354 CHAPTER 8. MORE TOPICS ON ELLIPTIC EQUATIONS

where we use the property that divcurluk = 0 so there is not derivative


hitting curluk when integrating by parts. Now since curluk is compact in
H −1 (Ω) , and pk converges weakly in H 1 (Ω) , we find that
Z Z Z
lim vk · wk ϕdx = −hcurlu, pDϕi = curlu · Dpϕdx = v · wϕdx .
k→∞ Ω Ω Ω

We will mimic this idea to prove Lemma 8.15.

Proof. Let wk ∈ H 2 (Ω) ∩ H01 (Ω) solve

−∆wk = vk in Ω,
wk = 0 on ∂Ω .

Then vk = curlcurlwk − Ddivwk , and kwk kH 2 (Ω) ≤ Ckvk kL2 (Ω) . Moreover,
let ϕ ∈ D(Ω) . Then

−∆(ϕcurlwk ) = −curlwk ∆ϕ − 2Dϕ · Dcurlwk + ϕcurlvk .

We claim that the right-hand side, at least for a subsequence, converges


strongly in H 1 (Ω)0 . The convergence of the last term is assumed at the
very beginning, while the first term converges strongly in L2 (Ω) by the
Rellich theorem. For the second term, by the definition of the dual space
norm,

kDϕ · Dcurl(wk − w)kH 1 (Ω)0 = sup hDϕ · Dcurl(wk − w), ψiH 1 (Ω)
kψkH 1 (Ω) =1

≤ 2kDϕkW 1,∞ (Ω) kcurl(wk − w)kL2 (Ω) → 0 as k → ∞ ,

where w denotes the limit of wk . By elliptic regularity,

kϕcurl(wk − w)kH 1 (Ω) → 0 as k → ∞ .

Therefore,
Z Z Z
uk · vk ϕdx = uk · curlcurlwk ϕdx − uk · Ddivwk , ϕdx
Ω ZΩ ΩZ

= uk · curl(ϕcurlwk )dx − uk · (Dϕ × curlwk )dx


ΩZ Z Ω
+ divuk divwk ϕdx + uk · Dϕdivuk dx .
Ω Ω
Z
It is easy to see that the right-hand side converges to u · vϕdx . 

8.5. COMPENSATED COMPACTNESS AND THE DIV-CURL LEMMA355

Example 8.16. Consider one dimensional Burgers equation ut + uux = 0


in R . Suppose u is the solution to the viscous Burgers equation

ut + u ux = uxx ∀x ∈ R,t > 0. (8.36)

We want to show that u converges to u in some proper sense. We consider


the two dimensional space R+ × R (treat t as another space dimension).
Then, (8.36) simply reads
(u )2
divt,x (u , ) = uxx . (8.37)
2
Multiplying (8.36) by u , we obtain
(u )3 (u )2 h (u )2 i
curlt,x (− , )= − (ux )2 . (8.38)
3 2 2 xx
Suppose that one knows that the right-hand side of (8.37) and (8.38) are
compact in H −1 , by the div-curl lemma,
(u )4 u1 u3 u22
− →− + in D0 (R) ,
12 3 4
where ui is the weak limits of (u )i . Therefore,

u4 = 4u1 u3 − 3u22 . (8.39)

Now consider the integral


Z
(u − u1 )4 dx .
R
Binomial expansion gives us
Z Z Z Z
(u − u1 ) dx = (u ) dx − 4 (u ) u1 dx + 6 (u )2 u21 dx
 4  4  3
R R Z RZ R

− 4 u u31 dx + u41 dx .
R R

Passing  to 0 , by (8.39) we find that


Z Z h i
 4
lim (u − u1 ) φdx = u4 − 4u1 u3 + 6u2 u21 − 4u41 + u41 φdx
→0 R R Z

= −3 (u2 − u21 )2 φdx .


R
Note that the left-hand side is always non-negative, while the right-hand side
is always non-positive. Therefore, u2 = u21 or more precisely,
 2
w. lim (u )2 = w. lim u .
→0 →0
356 CHAPTER 8. MORE TOPICS ON ELLIPTIC EQUATIONS

8.6 Exercises

Problem 8.1. Use the calculus on the boundary to show that


Z
2
kDukL2 (Ω) + H|u · N|2 dS = kcurluk2L2 (Ω) + kdivuk2L2 (Ω)
∂Ω
n−1
X Z h i
+ bαβ (u · ψ,α )(u · ψ,β ) + 2g αβ (u · ψ,α )(u · N),β dS ,
α,β=1 ∂Ω

n−1
where H again is the mean curvature of ∂Ω , and bαβ = − g αγ g βδ ψ,γδ ·N .
P
γ,δ=1
By the identity above, deduce that
h
kuk2H 1 (Ω) ≤ C(|∂Ω|H 2.5 ) kuk2L2 (Ω) + kdivuk2L2 (Ω) + kcurluk2L2 (Ω)
i
+ ku × Nk2H 0.5 (∂Ω) ∀ u ∈ H 1 (Ω) .

Also prove that


h
kuk2H k+1 (Ω) ≤ C(|∂Ω|H ` ) kuk2L2 (Ω) + kdivuk2H k (Ω) + kcurluk2H k (Ω)
i
+ k(∂u) × Nk2H k−0.5 (∂Ω) ∀ u ∈ H k+1 (Ω) .
(8.40)
where ` = max{2.5, k + 0.5} .

Problem 8.2. Prove Theorem 8.3 for the case n = 2 by the following steps:

1. Parameterize ∂Ω by (x(s), y(s)) , where s denote the arc-length, so that


the outward pointing normal is (y 0 (s), −x0 (s)) .

2. Solve (8.18). Because of (8.18b), there is a scalar function c defined


on ∂Ω so that

Dp(x(s), y(s)) = c(x(s))(x0 (s), y 0 (s)) .

Note that c can be zero at some points on ∂Ω .

3. Let B be a solution to
1
Z
−∆B = c(x)dS in Ω,
|Ω| ∂Ω
∂B
=c on ∂Ω ,
∂N
and then define v by the same formula used in the proof of Theorem
8.3. Note that now v is a scalar, not a vector.
8.6. EXERCISES 357

4. Show that vx = py and vy = −px .

5. The vector u := D⊥ v + Dp = (vy + px , −vx + py ) is an H01 (Ω)-solution


to the problem divu = f satisfying (8.17).

Problem 8.3. Let Ω ⊆ R3 be a bounded smooth domain, and u : Ω → R3 be


a vector-valued function. Suppose that u ∈ H 1 (Ω) and divu = 0 . Show that
u can be extended to the whole space so that the extension is still divergence-
free and u has compact support. In other words, find ũ ∈ H 1 (R3 ) such that
(1) divũ = 0 in R3 ; (2) ũ = u in Ω ; (3) spt(ũ) is compact.

Problem 8.4.

(1) Prove that the Leray projector P defined in Lemma ?? is linear.

(2) Prove that for u ∈ H 0 (Ω) , divP u = 0 .

Problem 8.5. Show that if u ∈ H k (Ω) is divergence free in Ω for some


k ∈ N , there exists a vector field w ∈ H k+1 (Ω) so that u = curlw and
kwkH k+1 (Ω) ≤ CkukH k (Ω) .
Chapter 9

Fluid dynamics

9.1 The equations

Let Ω ⊆ Rn be a smooth domain, occupied by a viscous incompressible fluid,


let u(x, t) denote the velocity vector field, and let p(x, t) denote the scalar
pressure function. The Navier-Stokes equations are defined as follows:

ut + (u · D)u + Dp = div(νDu) + f in Ω × (0, T ] , (9.1a)


divu = 0 in Ω × [0, T ] , (9.1b)
u=0 on ∂Ω × [0, T ] , (9.1c)
u = u0 on Ω × {t = 0} , (9.1d)

where the constant ν > 0 is the kinematic viscosity of the fluid (which, more
generally, may depend on x, t, and Du), f is the external forcing function,
and u0 is the initial velocity. The nonlinearity (u · D)u is the advection term
defined by

n
X ∂u
(u · D)u = uj .
∂xj
j=1

Note that (u · D)u is, in fact, the directional derivative of u in the direction
u , and is hence often expressed as Du u . Condition (9.1b) is often called
the divergence-free or the incompressibility condition, while the boundary
condition (9.1c) is known as the no-slip condition.
When the viscosity ν is large and diffusion dominates advection, the

358
9.2. INCOMPRESSIBLE EULER EQUATIONS ON A FIXED DOMAIN359

nonlinearity if ignored and (9.1) reduces to the linear Stokes equations:

ut − ν∆u + Dp = f in Ω × (0, T ] , (9.2a)


divu = 0 in Ω × [0, T ] , (9.2b)
u=0 on ∂Ω × [0, T ] , (9.2c)
u = u0 on Ω × {t = 0} . (9.2d)

On the other hand, when the viscosity ν is very small, and advection
dominates diffusion, we obtain the equations of a perfect or ideal incom-
pressible fluid, given by the Euler equations:

ut + (u · D)u + Dp = f in Ω × (0, T ] , (9.3a)


divu = 0 in Ω × [0, T ] , (9.3b)
u·N=0 on ∂Ω × [0, T ] , (9.3c)
u = u0 on Ω × {t = 0} . (9.3d)

9.2 Incompressible Euler equations on a fixed do-


main
9.2.1 Flow on a bounded and fixed domain

Let Ω denote a bounded, smooth, open subset of Rn with n = 2, 3, and let N


denote the outward-pointing unit normal to ∂Ω. We consider the following
system of equations describing the flow of an inviscid incompressible fluid:

ut + (u · D)u + Dp = 0 in Ω × (0, T ] , (9.4a)


divu = 0 in Ω × [0, T ] , (9.4b)
u·N=0 on ∂Ω × [0, T ] , (9.4c)
u = u0 on Ω × {t = 0} , (9.4d)

where the vector field u = (u1 , ..., un ) denotes the velocity of the fluid and
the scalar function p denotes the pressure in the fluid. The equation (9.4a)
denotes conservation of momentum, (9.4b) states that the fluid is incom-
pressible so that the fluid flow preserves volume, the boundary condition
(9.4c) states that fluid cannot leave the fluid container Ω, and (9.4d) pre-
scribes the initial velocity field u0 , which must also be divergence-free. The
ith component of the advection term can be expressed as [(u·D)u]i = ui ,j uj .
360 CHAPTER 9. Fluid dynamics

By first taking the divergence of (9.4a) and second computing its inner-
product with N on ∂Ω, we obtain the elliptic equation

−∆p = ui ,j uj ,i in Ω × [0, T ] , (9.5a)


∂p
= −ui ,j uj Ni on ∂Ω × [0, T ] . (9.5b)
∂N

An important observation is that the right-hand side of (9.5b) is, in fact,


quadratic in u. To see this, we rely on the identity (8.24), and write (in a
local chart (U, ψ) with n = N ◦ ψ)

−ui ,j ◦ ψ = −ui ,α g αβ ψ j ,β −ui ,n nj .

With τ α denoting g αβ ψ,β , it follows that

−(ui ,j uj Ni ) ◦ ψ = −(ui ,α τjα + ui ,n nj ) ni uj


= −ui ,α ni uj τjα = ui ni ,α (u · τ α )
= (uk τkβ ψ i ,β ) ni ,α (u · τ α )
= (u · τkβ )ψ i ,β g στ bασ ψ i ,τ (u · τ α ) = bαβ (u · τ α ) (u · τ β ) ;

thus (9.5b) can be written as

∂p
= b(u · τ, u · τ ) = b(u, u) . (9.5b’)
∂N

Standard elliptic estimates for the Neumann problem show that for s ≥ 0,
h i
kpkH s+1 (Ω) ≤ C kui ,j uj ,i kH s−1 (Ω) + kb(u, u)k 1 . (9.6)
H s− 2 (∂Ω)

9.2.2 The space of divergence-free vector fields on Ω

Lemma 9.1 (Normal trace lemma). Suppose that u ∈ L2 (Ω) and divu ∈
1
L2 (Ω). Then u · N ∈ H − 2 (∂Ω) and for a generic constant C > 0 depending
on Ω,
h i
ku · Nk 1
−2 ≤ C kukL2 (Ω) + kdivukL2 (Ω) .
H (∂Ω)

Proof. By density, assume that u ∈ C ∞ (Ω) and let φ be an arbitrary func-


tion in H 1 (Ω). According to Theorems 2.153 and 2.156, the trace operator
§9.2 Incompressible Euler equations 361

1
T : H 1 (Ω) → H 2 (∂Ω) is a continuous surjection so we may identify φ with
1
an H 2 (∂Ω) function. It follows that

ku · Nk 1
−2 = sup hu · N, φi ,
H (∂Ω) kφk =1
1
H 2 (∂Ω)

1 1
where h·, ·i denote the duality pairing between H 2 (∂Ω) and H − 2 (∂Ω). By
the divergence theorem,
Z

hu · N, φi = u · N φ dS

∂Ω
Z    
= divu φ + u · Dφ dx ≤ kdivukL2 (Ω) + kukL2 (Ω) kφkH 1 (Ω) .

Thanks to Lemma 9.1 we can make the following definition of an impor-


tant subspace of H s (Ω):

Definition 9.2 (Space of divergence-free vector fields with vanishing normal


trace). For k ≥ 0, set

V k = u ∈ H k (Ω) divu = 0 in Ω , u · N = 0 on ∂Ω ,


where divu is taken in the distributional sense. The space V k is a Hilbert


space when endowed with the H k (Ω)-norm.

The Hodge decomposition shows that when s ≥ 0,

H s (Ω) = V s ⊕L2 D[H s+1 (Ω)] ,

and there are natural projections onto the right-hand sides of this decom-
posion.

Definition 9.3 (Leray Projection). We define the projection operator P :


H s (Ω) → V s by P(w) = w −Dp, where p is the unique solution of the elliptic
equation
∂p
∆p = divw in Ω and = w · N on ∂Ω .
∂N
The divergence theorem shows that the solvability condition for this Neu-
mann problem is always satisfied.
362 CHAPTER 9. Fluid dynamics

The Leray projection allows us to reexpress the Euler equations as the


following system:
 
ut + P Du u = 0 in Ω × (0, T ] , (9.7a)
u = u0 on Ω × {t = 0} . (9.7b)

9.2.3 The Lagrangian flow of a velocity field

The passage to the so-called Lagrangian coordinates is extremely important


in the study of hyperbolic PDE.

Definition 9.4 (Lagrangian coordinates). We set η(x, t) to be the solution


of the ODE

∂t η(x, t) = u(η(x, t), t) for 0 < t ≤ T , and η = e on {t = 0} , (9.8)

where e(x) = x denotes the identity map on Ω. Given a particle labeled x at


t = 0, η(x, t) gives the position of this particle at time t ∈ (0, T ]. Whenever,
the velocity field u is Lipschitz continuous in x, uniformly in t, there exists
a unique solution η. The time T > 0 depends on u.

Lemma 9.5. If u · N = 0 on ∂Ω, then η(t)(∂Ω) ⊆ ∂Ω for t ∈ [0, T ].

Proof. Let x be any arbitrary and fixed element of ∂Ω. We must prove
that η(x, t) ∈ ∂Ω for all t ∈ [0, T ]. It suffices to show that each vector
ηt (x, t) is in the tangent space Tη(x,t) ∂Ω. This follows from the fact that
ηt (x, t) · N(η(x, t)) = [u · N](η(x, t)) = 0. 

Lemma 9.6. If divu = 0 in Ω × [0, T ], then det Dη = 1 in Ω × [0, T ].

Proof. Let A(x, t) = [Dη(x, t)]−1 denote the n × n inverse matrix, let J =
det Dη, and let
a(x, t) = J(x, t)A(x, t)

denote the matrix of cofactors. We compute the time-derivative of J as

Jt = asr ηtr ,s = JAsr ηtr ,s = J divu ◦ η ,

the last equality following from the chain rule. Since J(x, 0) = 1, given that
η(x, 0) = x, we see that J(x, t) = 1 whenever divu = 0. 
§9.2 Incompressible Euler equations 363

Proposition 9.7 (A polynomial-type inequality). For a constant M0 ≥ 0,


suppose that f (t) ≥ 0, t 7→ f (t) is continuous, and

f (t) ≤ M0 + C t P(f (t)) , (9.9)

where P denotes a polynomial function, and C is a generic constant. Then


for t taken sufficiently small (independently of the function f ≥ 0 satisfying
(9.9)), we have the bound
f (t) ≤ 2M0 .

We will see many applications of Proposition 9.7, which can be viewed


as a generalization of the nonlinear Gronwall inequality. In order to see the
first use of this result, it is convenient to first make the following
n
Definition 9.8 (The group of diffeomorphisms of Ω). For s > + 1, let
2
Ds (Ω) = ψ ∈ H s (Ω; Ω) ψ −1 exists, and ψ −1 ∈ H s (Ω; Ω) .


Ds (Ω) is the group of volume-preseving diffeomorphisms with group multi-


plication given by composition of maps on the right.
For each η ∈ Ds (Ω), the tangent space at η of this group is given by

Tη Ds (Ω) = v := u ◦ η u ∈ H s (Ω) and u · N = 0 on ∂Ω .




The tangent bundle T Ds (Ω) = ∪η∈Ds (Ω) Tη Ds (Ω) and is thus locally isomor-
phic to the space of pairs of “positions” and “velocities” (η, v).

Definition 9.9 (Subgroup of volume-preserving diffeomorphisms of Ω). For


n
s > + 1, we set
2
s
Dvol (Ω) = {η ∈ Ds (Ω) | det Dη = 1} .

s (Ω), the tangent space at η of this subgroup is given by


For each η ∈ Dvol

s
(Ω) = v := u ◦ η u ∈ V s .

Tη Dvol

Lemma 9.10. Suppose that u ∈ C([0, T1 ]; H 3 (Ω)) with u · N = 0 on ∂Ω ×


[0, T1 ]. Then there exists a unique solution η(x, t) of (9.8) with η ∈ C 1 ([0, T ]; Ds (Ω))
for some T ≤ T1 , and max kη(t)k2H 3 (Ω) ≤ 2kek2H 3 (Ω) .
[0,T ]
364 CHAPTER 9. Fluid dynamics

Proof. By the Sobolev embedding theorem, u is uniformly Lipschitz contin-


uous, so that there exists a unique solution η to (9.8). By the fundamental
theorem of calculus (with η(t) denoting η(·, t)),
Z t
η(t) = e + u ◦ η(s)ds ,
0

so that
X Z t
max kη(t)k2H 3 (Ω) ≤ kek2H 3 (Ω) + max kDα (u ◦ η)k2L2 (Ω) ds
[0,T ] [0,T ]
|α|≤3 0

≤ kek2H 3 (Ω) + T 2 max ku(t)k2H 3 (Ω) P max kη(t)k2H 3 (Ω) ,



[0,T ] [0,T ]

where we have used the chain rule, the Sobolev embedding theorem, and
Hölder’s and Jensen’s inequality to obtain the second inequality. Proposition
9.7 then provides the desired result for T > 0 taken small enough. 
n
Remark 9.11. If for s > + 1, u ∈ C([0, T1 ]; H s (Ω) with u · N = 0 on
2
∂Ω × [0, T1 ], then a generalization of the argument given in Lemma 9.10
shows that the flow η ∈ C 1 ([0, T ; Ds (Ω)) for T ≤ T1 . If u ∈ C([0, T1 ]; V s ),
then according to Lemma 9.6, each η ∈ C 1 ([0, T ]; Dvol
s (Ω)).

Remark 9.12. Since the flow map η(t) starts at the identity, for T >
0 taken sufficiently small, η(t) is a near-identity transformation, so that
max[0,T ] kη(t) − ek23 ≤ δ(T ) where δ(T ) → 0 as T → 0. Thus, by the inverse
function theorem η −1 (t) exists. Moreover, since Dη −1 ◦ η = [Dη]−1 , and
since [Dη]−1 = cof[Dη]/ det Dη, we see that each component of [Dη]−1 is a
n
polynomial function of partial derivatives of η. It follows that for s > + 1,
2
the map η 7→ η −1 : Ds → Ds is continuous.

Lemma 9.13. Let η ∈ D3 (Ω) and f ∈ H 3 (Ω). If kηkH 3 (Ω) ≤ C, then


kf ◦ ηkH 3 (Ω) ≤ Ckf kH 3 (Ω) .

Proof. The chain rule shows that

D(f ◦ η) = Df ◦ η Dη ,
D2 (f ◦ η) = D2 f ◦ η Dη Dη + Df ◦ η D2 η ,
D3 (f ◦ η) = D3 f ◦ η Dη Dη Dη + 3D2 f ◦ η D2 η Dη + Df ◦ η D3 η .
§9.2 Incompressible Euler equations 365

We then see that


Z
3
kD (f ◦ η)k2L2 (Ω) ≤ |D3 f ◦ η|2 |Dη(x)|6 dx
Ω Z
+ 3 |D2 f ◦ η|2 |D2 η(x)|2 |Dη(x)|2 dx
Ω Z
+ |Df ◦ η|2 |D3 η(x)|2 dx
Z Ω
6
≤ kDηkL∞ (Ω) |D3 f |2 | det Dη(x)|dx
Ω Z
2
+ 3kDf kL∞ (Ω) |D3 η(x)|2 dx

+ 3kDηk2L∞ (Ω) kD2 ηk2L6 (Ω) kD2 f ◦ ηk2L3 (Ω)
≤ P (kηkH 3 (Ω) )kf kH 3 (Ω) .

The lower-order terms are estimated in a similar fashion. 

9.2.4 Existence and uniqueness of solutions of the incom-


pressible Euler equations on fixed and bounded domain

We begin with the case that initial data has H 3 -class regularity.

Theorem 9.14. For u0 ∈ V 3 , there exists a unique solution u to (9.4) such


that

u ∈ C([0, T ]; V 3 ) ∩ C 1 ([0, T ]; V 2 ) ,

where T > 0 depends on ku0 kH 3 (Ω) .

Proof. Consider the subset of the Banach space C([0, T ], V 3 ) defined by

CT = {v ∈ C([0, T ]; H 3 (Ω) | v · N = 0 on ∂Ω × [0, T ] ,


max kv(t)k2H 3 (Ω) ≤ ku0 k2H 3 (Ω) + 1} ,
t∈[0,T ]

with T and M to be subsequently fixed. Choose v̄ ∈ CT , and set η̄ = e +


Rt 0 0 2
0 v̄(t )dt . The fundamental theorem of calculus shows that max[0,T ] kη̄(t)k3 ≤
C for T > 0 taken sufficiently small, and hence η̄(t) ∈ D3 (Ω) for each
t ∈ [0, T ] by the inverse function theorem.
366 CHAPTER 9. Fluid dynamics

Consider the linear problem

∂t ui + ui ,j ūj + pu ,i = 0 in Ω × (0, T ] , (9.10a)


divu = 0 in Ω × [0, T ] , (9.10b)
u·N=0 on ∂Ω × [0, T ] , (9.10c)
u = u0 on Ω × {t = 0} , (9.10d)

where the pressure function pu satisfies the linear elliptic equation

−∆pu = ui ,j ūj ,i in Ω × [0, T ] , (9.11a)


∂pu
= b(u, ū) on ∂Ω × [0, T ] , (9.11b)
∂N
and verifies the estimate

kpu kH 4 (Ω) ≤ C kui ,j ūj ,i kH 2 (Ω) + kb(u, ū)kH 2.5 (∂Ω)




≤ CkukH 3 (Ω) kūkH 3 (Ω) . (9.12)

If we define the Lagrangian velocity v associated to the solution u of (9.10)


by v := u ◦ η̄, then (9.10a) can be written as

vt + Dpu (η̄) = 0 . (9.10a’)

so that by the fundamental theorem of calculus,


Z t
v(t) = u0 − Dpu (η̄)ds . (9.10a”)
0

For v̄ ∈ CT fixed, we next construct a solution to (9.10a’) as the fixed-


point of the map v 7→ ṽ , where v ∈ CT is given and
Z t
ṽ(t) = u0 − Dpu (η̄)ds .
0
The chain rule, the Sobolev embedding theorem, and Hölder’s and Jensen’s
inequality show that

max kṽ(t)k2H 3 (Ω) ≤ ku0 k2H 3 (Ω) + T max kpu k2H 4 (Ω) P ( max kη̄(t)k2H 3 (Ω) )
t∈[0,T ] t∈[0,T ] t∈[0,T ]

According to Lemma 9.10, maxt∈[0,T ] kη̄(t)k2H 3 (Ω) ≤ C, so that with (9.12),

max kṽ(t)k2H 3 (Ω) ≤ ku0 k2H 3 (Ω) + C T max ku(t)k2H 3 (Ω)


t∈[0,T ] t∈[0,T ]

≤ ku0 k2H 3 (Ω) + C T max kv(t)k2H 3 (Ω) ≤ ku0 k2H 3 (Ω) + C T ,


t∈[0,T ]
§9.2 Incompressible Euler equations 367

so that for T > 0 sufficiently small, ṽ ∈ CT . Next, let v (a) ∈ CT for a = 1, 2,


(a)
define u(a) := v (a) ◦ η̄ −1 , and let pu denote the solution to

−∆p(a) (a) i j
u = [u ] ,j ū ,i in Ω × [0, T ] ,
(a)
∂pu
= b(u(a) , ū) on ∂Ω × [0, T ] ,
∂N
and Z t
(a)
ṽ (t) = u0 − Dp(a)
u (η̄)ds .
0
Then, following our estimates above, we see that

max kṽ (1) (t) − ṽ (2) (t)k ≤ C T max kv (1) (t) − v (2) (t)k .
t∈[0,T ] t∈[0,T ]

The contraction mapping principle provides a unique fixed-point to the map


v 7→ ṽ in CT which is a solution to (9.10a”), and hence (9.10a) and (9.10d)
are satisfied. The conditions (9.10b) and (9.10c) are then verified by the
definition of pu . Having obtained a solution u ∈ V 3 to the linear problem
(9.10), we can now construct a solution to the incompressible Euler equations
(9.4) as the fixed-point of the map v̄ 7→ v in the space CT for T > 0 taken
sufficiently small.
From (9.10a”), we once again see that

max kv(t)k2H 3 (Ω) ≤ ku0 k2H 3 (Ω) + C T max kv(t)k2H 3 (Ω) ,


t∈[0,T ] t∈[0,T ]

so that v ∈ CT for T > 0 small enough.


Next, suppose that v̄1 and v̄2 are both in CT , and for a = 1, 2, set
Rt
η̄a (t) = e + 0 v̄a , ūa = v̄a ◦ η̄a−1 , ua = va ◦ η̄a−1 ,and let Pa := paua solve

−∆paua = uia ,j ūja ,i in Ω × [0, T ] , (9.14a)


∂paua
= b(ua , ūa ) on ∂Ω × [0, T ] , (9.14b)
∂N
It follows that with
Z t Z t
v1 (t) = u0 − DP1 (η̄1 )ds and v2 (t) = u0 − DP2 (η̄2 )ds ,
0 0

we have the estimate

max kv1 (t) − v2 (t)k2H 3 (Ω) ≤ T 2 max kDP1 ◦ η̄1 − DP2 ◦ η̄2 k2H 3 (Ω) .
t∈[0,T ] t∈[0,T ]
368 CHAPTER 9. Fluid dynamics

It remains to show that



kDP1 ◦ η̄1 − DP2 ◦ η̄2 kH 3 (Ω) ≤ C kv1 − v2 kH 3 (Ω) + kv̄1 − v̄2 kH 3 (Ω) , (9.15)

as the contraction mapping principle would then conclude the proof.


Setting ξ = η̄1 ◦ η̄2−1 , Lemma 9.13 shows that

kDP1 ◦ η̄1 − DP2 ◦ η̄2 kH 3 (Ω) ≤ CkDP1 ◦ ξ − DP2 kH 3 (Ω) .

In order to estimate this right-hand side, we will rely on the elliptic estimate
(8.16) and show that

kdiv(DP1 ◦ ξ − DP2 )kH 2 (Ω) ≤ C kv1 − v2 kH 3 (Ω) + kv̄1 − v̄2 kH 3 (Ω) ,
(9.16)

k curl(DP1 ◦ ξ − DP2 )kH 2 (Ω) ≤ C kv1 − v2 kH 3 (Ω) + kv̄1 − v̄2 kH 3 (Ω) ,
(9.17)

k(DP1 ◦ ξ − DP2 ) · NkH 2.5 (∂Ω) ≤ C kv1 − v2 kH 3 (Ω) + kv̄1 − v̄2 kH 3 (Ω) ,
(9.18)

and

kDP1 ◦ ξ − DP2 kL2 (Ω) ≤ C kv1 − v2 kH 3 (Ω) + kv̄1 − v̄2 kH 3 (Ω) . (9.19)

For this purpose, we write

div(DP1 ◦ ξ) = ∆P1 ◦ ξ + P1 ,ij ◦ξ (ξ j ,i −δij ) ,

and we rewrite (9.14) as


h i
−∆Pa = vai ,r [Āa ]rj v̄aj ,s [Āa ]si ◦ η̄a−1 in Ω × [0, T ] , (9.20a)
∂Pa
= b(va , v̄a ) ◦ η̄a−1 on ∂Ω × [0, T ] . (9.20b)
∂N

Hence,

kdiv(DP1 ◦ ξ − DP2 )kH 2 (Ω)


≤ k∆P1 ◦ ξ − ∆P2 kH 2 (Ω) + kP1 ,ij ◦ξ(ξ j ,i −δij )kH 2 (Ω)
 
≤ C k∆P1 ◦ η̄1 − ∆P2 ◦ η̄2 kH 2 (Ω) + kP1 ,ij ◦ξkH 2 (Ω) k(η̄1j ,i −η̄2j ,i )kH 2 (Ω)
 
≤ C kv1i ,r [Ā1 ]rj v̄1j ,s [Ā1 ]si − v2i ,r [Ā2 ]rj v̄2j ,s [Ā2 ]si kH 2 (Ω) + k(η̄1j ,i −η̄2j ,i )kH 2 (Ω) ,
9.3. THE STOKES EQUATIONS 369

the last inequality following from (9.20a). It follows that (9.16) holds; sim-
ilarly, (9.18) holds by replacing the divergence operator with the normal
trace, and using (9.20b). The inequality (9.17) follows from the fact that
curl DP1 = 0.
In order to obtain the inequality (9.19), we see that

kDP1 ◦ ξ − DP2 kL2 (Ω) ≤ CkDP1 ◦ η̄1 − DP2 ◦ η̄2 kL2 (Ω) .

With Qa = Pa ◦ η̄a for a = 1, 2, we write (9.20) as


 
−[Aa ]ji [Aa ]ki Qa ,k ,j = vai ,r [Āa ]rj v̄aj ,s [Āa ]si in Ω × [0, T ] , (9.21a)
Qa ,k [Aa ]ki [Aa ]ji Nj = b(va , v̄a ) on ∂Ω × [0, T ] . (9.21b)

Since P1 ,i ◦η̄1 −P2 ,i ◦η̄2 = Q1 ,k [A1 ]ki −Q2 ,k [A2 ]ki , elliptic estimates for (9.21)
show that (9.19) holds. 

An induction argument show that the more general statement of the


theorem holds:

Theorem 9.15 (Existence and uniqueness for the incompressible Euler


n
equations). For u0 ∈ V s , s > + 1, there exists a unique solution u to
2
(9.4) such that
u ∈ C([0, T ]; V s ) ∩ C 1 ([0, T ]; V s−1 ) ,

where T > 0 depends on ku0 kH s (Ω) .

9.3 The Stokes equations


9.3.1 The steady state case

In this section, we are concerned with the steady state Stokes equations
(with ν = 1)

−∆u + Dp = f in Ω × (0, T ) , (9.22a)


divu = 0 in Ω × (0, T ) , (9.22b)
u=0 on ∂Ω × (0, T ) . (9.22c)

Note that since u and p are time independent, we need the initial condition
(9.2d) no more.
370 CHAPTER 9. Fluid dynamics

Similar to V k defined in the previous section, for k ∈ N we define the


space W k by
n o
W k ≡ u ∈ H01 (Ω; Rn ) ∩ H k (Ω; Rn ) divu = 0

with the usual H k (Ω)-norm. W k is a Banach space for all k ∈ N.


Suppose that ϕ ∈ W 1 , then (Dp, ϕ)L2 (Ω) = 0 if p for p ∈ H 1 (Ω). This
observation leads to the following definition.

Definition 9.16. A vector valued function u ∈ W 1 is said to be a weak


solution to (9.22) if

(Du, Dϕ)L2 (Ω;Rn ) = hf, ϕiW 1 ∀ ϕ ∈ W 1. (9.23)

It is easy to see from the Lax-Milgram Theorem that a unique weak


solution to (9.22) exists. Moreover, the weak solution u satisfies

kukH01 (Ω) = kDukL2 (Ω) ≤ kf kH −1 (Ω) .

The question of how to recover the pressure function is given by the following

Lemma 9.17 (Lagrange multiplier lemma). Suppose that T : H01 (Ω; Rn ) →


R is a bounded linear functional such that T ϕ = 0 for all ϕ ∈ W 1 . Then
there exists p ∈ L2 (Ω)/R, unique modulo a constant, such that

T ϕ = (p, divϕ)L2 (Ω) ∀ ϕ ∈ H01 (Ω; Rn ).

Moreover,

kpkL2 (Ω) ≤ kT kB(H01 (Ω;Rn ),R) . (9.24)

Note, that we are using k · kB(H01 (Ω;Rn ),R) to denote the operator norm.

Proof. Given p ∈ L2 (Ω)/R, define a linear functional Tp : H01 (Ω; Rn ) → R


by Tp ϕ = (p, divϕ)L2 (Ω) . Then

kTp kB(H01 (Ω;Rn ),R) = sup |(p, divϕ)L2 (Ω) | ≤ kpkL2 (Ω) ;
kDϕkL2 (Ω) =1

thus Tp is bounded. By the Riesz representation theorem, there exists Qp ∈


H01 (Ω) such that

Tp ϕ = (p, divϕ)L2 (Ω) = (Qp, ϕ)H01 (Ω) ,


§9.3 The Stokes equations 371

and kQpkH01 (Ω) = kTp kB(H01 (Ω;Rn ),R) ≤ CkpkL2 (Ω) . Therefore, Q : L2 (Ω)/R →
H01 (Ω) is a bounded linear functional.
On the other hand, for this given p ∈ L2 (Ω)/R, by Theorem 8.3 there
exists ϕ ∈ H01 (Ω) such that divϕ = p. Therefore,

Tp ϕ = kpk2L2 (Ω) ≤ kQpkH01 (Ω) kpkH01 (Ω)

which implies that

kQpkH01 (Ω) ≥ kpkL2 (Ω) . (9.25)

As a consequence, R(Q), the range of Q, is closed (see the proof of the


Lax-Milgram Theorem for the detail). Since R(Q)⊥ = W 1 ,

H01 (Ω) = R(Q) ⊕H01 (Ω) R(Q)⊥ = R(Q) ⊕H01 (Ω) W 1 .

Since T : H01 (Ω; Rn ) → R is bounded, by the Riesz representation theorem


there exists ψ ∈ H01 (Ω; Rn ) such that T ϕ = (ψ, ϕ)H01 (Ω) and kψkH01 (Ω;Rn ) =
kT kB(H01 (Ω;Rn ),R) . By the decomposition above, ψ = ψ1 + ψ2 for some ψ1 ∈
R(Q) and ψ2 ∈ W 1 . Let ψ1 = Qp. For ϕ ∈ H01 (Ω), write ϕ = ϕ1 + ϕ2 for
some ϕ1 ∈ R(Q) and ϕ2 ∈ W 1 . Then

T ϕ = T ϕ1 = (ψ, ϕ1 )H01 (Ω;Rn ) = (ψ1 , ϕ1 )H01 (Ω;Rn ) = (ψ1 , ϕ)H01 (Ω;Rn )


= (Qp, ϕ)H01 (Ω;Rn ) = (p, divϕ)L2 (Ω) .

Moreover, (9.25) implies that

kpkL2 (Ω) ≤ kQpkH01 (Ω) = kψ1 kH01 (Ω) ≤ kψkH01 (Ω) = kT kB(H01 (Ω;Rn ),R) .


Given f ∈ H −1 (Ω; Rn ), let u be the weak solution to (9.22). Define


T : H01 (Ω; Rn ) → R by T ϕ = (Du, Dϕ)L2 (Ω) − hf, ϕiH01 (Ω;Rn ) . Then

kT kB(H01 (Ω;Rn ),R) = sup |T ϕ| ≤ kDukL2 (Ω) + kf kH −1 (Ω) ≤ 2kf kH −1 (Ω)


kDϕkL2 (Ω) =1

implies that T : H01 (Ω; Rn ) → R is a bounded linear functional. Moreover,


T ϕ = 0 for all ϕ ∈ W 1 . Therefore, Lemma 9.17 provides the existence of a
pressure p ∈ L2 (Ω) such that

(Du, Dϕ)L2 (Ω) − hf, ϕiH01 (Ω;Rn ) = (p, divϕ)L2 (Ω) ∀ ϕ ∈ H01 (Ω; Rn ),
372 CHAPTER 9. Fluid dynamics

and this p satisfies

kpkL2 (Ω) ≤ kT kB(H01 (Ω;Rn ),R) ≤ 2kf kH −1 (Ω) . (9.26)

Based on the discussion above, we have the following alternative definition


of weak solution.

Definition 9.18. The pair (u, p) ∈ W 1 ×L2 (Ω) is said to be a weak solution
to (9.22) if

(Du, Dϕ)L2 (Ω) + (p, divϕ)L2 (Ω) = hf, ϕiH01 (Ω;Rn ) ∀ ϕ ∈ H01 (Ω; Rn ). (9.27)

Regularity theory for the Stokes problem

Now suppose that f ∈ L2 (Ω). Similar to the regularity theory of Section


5.9, we expect that inversion of the Stokes operator provides a two-derivative
gain of regularity with respect to the forcing function f . In particular, for
f ∈ L2 (Ω), we should be able to show that u ∈ W 2 .
Indeed, this is the case, and the proof that u ∈ W 2 is somewhat similar
to the proof of elliptic regularity in Section 5.9, once we agree to use the
variational formulation (9.27) rather then (9.23). The primary advantage
of the form (9.27) comes from the removal of the divergence-free constraint
on the test functions. Notice that if we set v = η ∗ ζ 2 D2 (η ∗ u) as we
 

introduced in Section 5.9 for the proof of interior regularity, then the test
function v ∈ H01 (Ω; Rn ) but the presence of the cut-off function does not
preserve the divergence-free constraint and divv 6= 0.

Theorem 9.19. Suppose that ∂Ω is of class C m+1 for some m ∈ N. Then


for any f ∈ H m−1 (Ω), there exists a unique solution (u, p) ∈ W m+1 ×
H m (Ω)/R to the steady Stokes problem

−∆u + Dp = f in Ω × (0, T ) ,
divu = 0 in Ω × (0, T ) ,
u=0 on ∂Ω × (0, T ) .

Moreover, (u, p) satisfies

kukH m+1 (Ω) + kpkH m (Ω) ≤ Ckf kH m−1 (Ω) . (9.29)


§9.3 The Stokes equations 373

Proof. We first prove that u ∈ W 2 ; the general case follows by an induction


argument.
For the proof, we will use the same cut-off functions ζ, ζm , and local
charts θm as we introduced in Section 5.9.
Step 1. Interior estimates.
Let v = η ∗ ζ 2 (η ∗ u),k ,k . The use of v as a test function in (9.27)
 

implies that

− (Du, Dv)L2 (Ω) = −(f, v)L2 (Ω) + (p, divv)L2 (Ω)


≤ Cδ kf k2L2 (Ω) + δkζD2 (η ∗ u)k2L2 (Ω) − (p, divv)L2 (Ω) . (9.30)

Since divu = 0,
Z
p η ∗ [ζ 2 (η ∗ ui ),k ],k ,i dx

(p, divv)L2 (Ω) =

Z
p η ∗ [ζζ,i (η ∗ ui ),k ],k dx;

=2 (9.31)

thus by (9.26) and Young’s inequality,


h i
(p, divv)L2 (Ω) ≤ CkpkL2 (Ω) kukH 1 (Ω) + kζD2 (η ∗ u)kL2 (Ω)
≤ Cδ kf k2L2 (Ω) + δkζD2 (η ∗ u)k2L2 (Ω) . (9.32)

On the other hand, just as in Section 5.9, we have that

−(Du,Dv)L2 (Ω) = D(η ∗ u),k , D[(ζ 2 (η ∗ u),k ] L2 (Ω)




≥ kζD2 (η ∗ u)k2L2 (Ω) − CkD(η ∗ u)kL2 (Ω) kζD2 (η ∗ u)k2L2 (Ω)
≥ (1 − δ)kζD2 (η ∗ u)k2L2 (Ω) − Cδ kuk2H 1 (Ω)
≥ (1 − δ)kζD2 (η ∗ u)k2L2 (Ω) − Cδ kf k2L2 (Ω) .

Therefore, by choosing δ > 0 small enough, we conclude that

kζD2 (η ∗ u)k2L2 (Ω) ≤ Ckf kL2 (Ω)

for some constant C = C(kDζkL∞ (Ω) ). Since the bound on the right-hand
side is independent of , we may pass to a weak limit (of a subsequence) and
obtain that

kζD2 uk2L2 (Ω) ≤ Ckf kL2 (Ω) . (9.33)


374 CHAPTER 9. Fluid dynamics

2 (Ω).
We note that estimate (9.33) implies that u ∈ Hloc
Given ψ ∈ H 2 (Ω), let ϕ = ζDψ in (9.27). With Defu denoting the
1
symmetric part of Du; that is, Defu = Du + (Du)T , we find that

2

(ζp, ∆ψ)L2 (Ω) = (ζf − pDζ, Dψ)L2 (Ω) − Du, D(ζDψ) L2 (Ω)

= (ζf − pDζ − DefuDζ, Dψ)L2 (Ω) . (9.34)

Identity (9.34) is the same as saying that q = ζp is a distributional solution


to the elliptic equation

−∆q = div(ζf − pDζ − DefuDζ) in Ω, (9.35a)


q=0 on ∂Ω . (9.35b)

However, since div(ζf − pDζ − DefuDζ) ∈ H −1 (Ω), there exists a weak


solution q̃ ∈ H01 (Ω) to (9.35). Moreover, since (q − q̃, ∆ψ)L2 (Ω) = 0 for all
ψ ∈ C02 (Ω), q̃ must agree with q. Therefore, ζp ∈ H01 (Ω) satisfying

kζpkH 1 (Ω) ≤ Ckdiv(ζf − pDζ − DefuDζ)kH −1 (Ω) ≤ Ckf kL2 (Ω) (9.36)

for some constant C = C(kDζkL∞ (Ω) ), and estimate (9.36) implies that
1 (Ω).
p ∈ Hloc
2 (Ω) and p ∈ H 1 (Ω), we can integrate by
Having established u ∈ Hloc loc
parts (to move the derivative on the test function back onto u and p in
(9.23)) and obtain that (f + ∆u − Dp, ϕ)L2 (Ω) = 0 for all ϕ ∈ C01 (Ω). This
implies that

−∆u + Dp = f a.e. in Ω. (9.37)

Step 2. Boundary estimates for u ◦ θm .


Define A = (Dθm )−1 , ξm = ζm ◦ θm and um = u ◦ θm as in the proof of
  −1
the elliptic regularity. Let v i = θm,`
 i 2 Λ (A` umj ),
Λ  ξm  j α ,α ◦ θm be a test

function in (9.27). Using the commutator notation,

 
A, B u = A B u − B A u ,
§9.3 The Stokes equations 375

we have that

(v ◦ θm )i = θm,`
i 2
Λ (A`j umj ),α ,α

Λ ξm
i 2
Λ (A`j ,α umj ) ,α +2θm,`
i
Λ ξm ξm ,α Λ (A`j umj ,α )
 
= θm,` Λ  ξm
i 2
∂α Λ , A`j umj ,α + θm,`i 2 `
Aj ,α Λ umj ,α )
  
+ θm,` Λ  ξm Λ (ξm
i 2 `
Aj (Λ umj ),αα .

+ θm,` Λ  ξm

Applying the commutation estimate (5.46) to estimate the third term on


the right-hand side, we find that

h ` 2 i
mj

kvkL2 (Ω) ≤ C kukH 1 (Ω) + Aj ξm (Λ u
,αα L2 (B+ (0,rm ))

h i
¯
≤ C kf kL2 (Ω) + kξm ∂DΛ  um
k L (B+ (0,rm )) ,
2


where C = C kDζm kL∞ (B+ (0,rm )) , kDθm kL∞ (B+ (0,rm )) , kAkW 1,∞ (B+ (0,rm )) .
As a consequence, similar to (9.30), we obtain that

¯
−(Du, Dv)L2 (Ω) ≤ Cδ kf k2L2 (Ω) + δkξm ∂DΛ m 2
 u kL2 (B+ (0,rm )) + (p, divv)L2 (Ω) .

To estimate (p, divv)L2 (Ω) , we note that the change of variable x 7→ θm (x)
implies that

Z h
i 2
  −1 i
(p, divv)L2 (Ω) = p θm,` Λ ξm Λ (A`j umj ),α ,α ◦ θm ,i dx
ZΩ
(p ◦ θm )Aki θm,`
 i 2
Λ (A`j umj ),α ,α ,k dx
 
= Λ  ξm
B+ (0,rm )
Z
2
Λ (Akj umj ),α ,kα dx,

= (p ◦ θm )Λ ξm
B+ (0,rm )

where we have used that Aki θm,`


i = δ`k together with the Piola identity Aki ,k =
0 (due to the assumption that det(Dθm ) = 1) to conclude the last equality.
376 CHAPTER 9. Fluid dynamics

Since divu = 0, Akj umj ,k = (divu) ◦ θm = 0 in B+ (0, rm ). Therefore,


Z
(p ◦ θm )Λ ξm ξm ,k Λ (Akj umj ),α ,α dx

(p, divv)L2 (Ω) = 2
B (0,r )
Z + m
(p ◦ θm )Λ ξm ξm ,k Λ (Akj ,α umj ) ,α dx

=2
B+ (0,rm )
Z
(p ◦ θm )Λ (ξm ξm ,k ),α Λ (Akj umj ,α ) dx

+2
B (0,r )
Z + m
(p ◦ θm )Λ ξm ξm ,k ∂α ( Λ , Akj umj ,α ) dx
  
+2
B (0,r )
Z + m
(p ◦ θm )Λ ξm ξm ,k Akj ,α (Λ umj ),α ) dx

+2
B (0,r )
Z + m
(p ◦ θm )Λ ξm ξm ,k Akj (Λ umj ),αα dx.

+2
B+ (0,rm )

Using (5.46) again to estimate the third term on the right-hand side, by
(9.26) and Young’s inequality we find that
h i
¯
(p, divv)L2 (Ω) ≤ CkpkL2 (Ω) kukH 1 (Ω) + kξm ∂DΛ  um
k 2
L (Ω)
¯
≤ Cδ kf k2L2 (Ω) + δkξm ∂DΛ m 2
 u kL2 (Ω) ; (9.38)

thus

¯
−(Du, Dv)L2 (Ω) ≤ Cδ kf k2L2 (Ω) + δkξm ∂DΛ m 2
 u kL2 (B+ (0,rm )) .

Now we turn our attention to the term which contributes the energy.
First the change of variable x 7→ θm (x) implies that
Z
(Du, Dv)L2 (Ω) = Ark umi ,r Ask v i ,s dx
B+ (0,rm )
Z h  i
= Ark umi ,r Ask θm,` i 2
Λ ξm Λ (A`j umj ),α ,α ,s dx.
B+ (0,rm )

Integrating by parts in xα , by the property of convolution we obtain that


Z
2
− (Du, Dv)L2 (Ω) = ξm Λ (Ark Ask θm,`
i
umi ,r ),α Λ (A`j umj ,s ),α dx
B+ (0,rm )
Z h i
2
+ ξm Λ (Ark Ask θm,`
i
umi ,r )Λ (A`j umj ,α ),α +R1 dx,
B+ (0,rm )

where R1 consists of products of first derivatives of um , and satisfies


Z
|R1 |dx ≤ Ckuk2H 1 (Ω) ≤ Ckf k2L2 (Ω) .
B+ (0,rm )
§9.3 The Stokes equations 377

Since Λ (A`j umj ,α ),α = ∂α A`j (Λ umj ),α + ∂α Λ , A`j umj ,α , by (5.46),
  

Z
2
ξm Λ (Ark Ask θm,`
i
umi ,r )Λ (A`j umj ,α ),α dx
B+ (0,rm )
h i
≤ CkukH 1 (Ω) kukH 1 (Ω) + kξm ∂DΛ ¯ m
 u kL2 (B+ (0,rm ))

¯
≤ Cδ kf k2L2 (Ω) + δkξm ∂DΛum 2
kL2 (B+ (0,rm )) .

Moreover,

Λ (Ark Ask θm,`


i
umi ,r ),α Λ (A`j umj ,s ),α
h i
= Ark Ask θm,`i i
(Λ umi ),rα +∂α Λ , Ark Ask θm,`
 mi i
u ,r +(Ark Ask θm,` ),α (Λ umi ),r

h i
× A`j (Λ umj ),sα +∂α Λ , A`j umj ,s +A`j ,α (Λ umj ),s ;
 

thus by (5.46) again,


Z
2
ξm i
Λ (Ark Ask θm,` umi ,r ),α Λ (A`j umj ,s ),α dx
B+ (0,rm )
Z
2 r s
≥ ξm Ak Ak (Λ umj ),rα (Λ umj ),sα dx − Ckuk2H 1 (Ω)
B+ (0,rm )
¯
− CkukH 1 (Ω) kξm ∂DΛum
kL2 (B+ (0,rm ))
¯
≥ (λ − δ)kξm ∂DΛum 2
kL2 (B+ (0,rm )) − Cδ kf k2L2 (Ω) ,

here we use the fact that Ark Ask is positive definite whose smallest eigen-
value is bigger than some positive constant λ, as in the proof of the elliptic
regularity. Therefore, by choosing δ > 0 small enough, we conclude that

¯
kξm ∂DΛ m
 u kL2 (B+ (0,rm )) ≤ Ckf kL2 (Ω) ,


where C = C kDζm kL∞ (B+ (0,rm )) , kDθm kL∞ (B+ (0,rm )) , kAkW 1,∞ (B+ (0,rm )) .
The bound on the right-hand side is independent of , so we may pass to
the weak limit (of a subsequence) and obtain that

¯
kξm ∂D(u ◦ θm )kL2 (B+ (0,rm )) ≤ Ckf kL2 (Ω) , (9.39)

Step 3. Boundary estimates for p ◦ θm .


We first assume that Ω = Rn+ to illustrate the idea. Letting ϕ = ψ,α in
(9.27), we find that

(Du, Dψ,α )L2 (Ω) + (p, divψ,α )L2 (Ω) = (f, ψ,α )L2 (Ω) .
378 CHAPTER 9. Fluid dynamics

Integrating by parts in xα , we find that

(p,α , divψ)L2 (Ω) = −(f, ψ,α )L2 (Ω) − (Du,α , Dψ)L2 (Ω) . (9.40)

Define a bounded functional T : H01 (Ω) → R by

T (ψ) = −(f, ψ,α )L2 (Ω) − (Du,α , Dψ)L2 (Ω) .

By (9.40), we know that T (ψ) = 0 for all ψ ∈ W 1 ; thus the Lagrange multi-
2
Z r ∈ L (Ω)/R such that T (ψ) = (r, divψ)L2 (Ω) .
plier lemma provides a unique
On the other hand, since p,α dx = 0, r = p,α ; thus

h i
¯
kp,α kL2 (Ω) ≤ CkT kB(H01 (Ω;Rn ),R) ≤ C kf kL2 (Ω) + k∂Duk L (Ω) ≤ Ckf kL2 (Ω) .
2

Now suppose that Ω is a general bounded C 2 -domain. Define pm = p◦θm


 −1
and f m = f ◦ θm . Let ϕi = θm,`
i ξm Λ (A`j ψ j ),α ◦ θm be a test function in
(9.27) for some ψ ∈ H01 (Ω; Rn ). Making a change of variable x 7→ θm (x), by
Aki θm,`
i = δ`k and the Piola identity Aki ,k = 0 we find that

− pm , (ξm Λ (Akj ψ j ),α ),k L2 (Ω) = − pm , Aki (θm,`i


ξm Λ (A`j ψ j ),α ),k L2 (B+ (0,rm ))
 

= Ajr umi ,j , Akr (θm,`


i
ξm Λ (A`j ψ j ),α ),k L2 (B+ (0,rm ))


− f m , θm,`
i
ξm Λ (A`j ψ j ),α L2 (B+ (0,rm )) .


Commuting derivatives and convolution operators about multiplication with


the A matrices, the identity above can be written as

Λ (ξm pm ),α , Akj ψ j ,k L2 (Ω) − Λ (ξm ,k pm ), (Akj ψ j ),α L2 (Ω)


 

= Λ (Akr Ajr (θm,`i


ξm ),k umi ),j , (A`j ψ j ),α L2 (B+ (0,rm ))


− A`j ,k Λ (Akr Ajr θm,`


i
ξm umi ,j ),α , ψ j L2 (B+ (0,rm ))

(9.41)
− A`j Λ (Akr Ajr θm,`i
ξm umi ,j ),α , ψ j ,k L2 (B+ (0,rm ))


i
ξm f m ), (A`j ψ j ),α L2 (B+ (0,rm )) .

− Λ (θm,`

The identity above motivates us to define the linear functional T given by

T (ϕ) = Λ (ξm ,k pm ), (Akj ψ j ),α L2 (Ω)




+ Λ (Akr Ajr (θm,`


i
ξm ),k umi ),j , (A`j ψ j ),α L2 (B+ (0,rm ))


− A`j ,k Λ (Akr Ajr θm,`


i
ξm umi ,j ),α , ψ j L2 (B+ (0,rm ))


− A`j Λ (Akr Ajr θm,`


i
ξm umi ,j ),α , ψ j ,k L2 (B+ (0,rm ))


i
ξm f m ), (A`j ψ j ),α L2 (B+ (0,rm )) ,

− Λ (θm,`
§9.3 The Stokes equations 379

where ψ on the right-hand side is the same as ϕ ◦ θm . We note that T :


H01 (Ω; Rn ) → R is bounded since

kT kB(H01 (Ω;Rn ),R)


h i
¯ m kL2 (Ω) + kf kL2 (Ω) ≤ Ckf kL2 (Ω)
≤ C kpkL2 (Ω) + kukH 1 (Ω) + kξm ∂Du

for some constant C = C kθm kW 2,∞ (B+ (0,rm )) , kAkW 1,∞ (B+ (0,rm )) . More-
over, T ϕ = 0 for all ϕ ∈ W 1 . Therefore, by the Lagrange multiplier lemma,
there exists a unique r ∈ L2 (Ω)/R such that T (ϕ) = (r, divϕ)L2 (Ω) , and

krkL2 (Ω) ≤ kT kB(H01 (Ω;Rn ),R) ≤ Ckf kL2 (Ω) . (9.42)

On the other hand, by the property of the convolution,


Z Z Z
m −1 m
Λ (ξm p ),α ◦ θm dx = Λ (ξm p ),α dx = (ξm pm ),α dx = 0
Ω B+ (0,rm ) B+ (0,rm )

−1 ∈ L2 (Ω)/R. Since for ϕ = ψ ◦ θ −1 ,


which implies that Λ (ξm pm ),α ◦ θm m

−1
Λ (ξm pm ),α , Aki ψ i ,k = Λ (ξm pm ),α ◦ θm
 
L2 (Ω)
, divϕ L2 (Ω) ,

the unique existence of r and (9.41) then imply that

−1
r = Λ (ξm pm ),α ◦ θm .

Therefore, (9.42) implies that


by det(Dθm ) = 1 −1
kΛ (ξm pm ),α kL2 (Ω) = kΛ (ξm pm ),α ◦ θm kL2 (Ω) ≤ Ckf kL2 (Ω) .

Since the bound on the right-hand side is independent of , we may pass to


the limit and obtain that

¯ ◦ θm )kL2 (B (0,r )) ≤ Ckf kL2 (Ω) .


kξm ∂(p (9.43)
+ m

Step 4. The estimates for ξm ∂N DΛ (u ◦ θm ) and ξm ∂N (p ◦ θm ).


We first assume that Ω = R3+ to illustrate the idea. Note that (9.37)
implies that

−u1 ,11 −u1 ,22 −u1 ,33 +p,1 = f 1 a.e. in Ω, (9.44a)


−u2 ,11 −u2 ,22 −u2 ,33 +p,2 = f 2 a.e. in Ω, (9.44b)
−u3 ,11 −u3 ,22 −u3 ,33 +p,3 = f 3 a.e. in Ω. (9.44c)
380 CHAPTER 9. Fluid dynamics

In (9.44), four functions u1 ,33 , u2 ,33 , u3 ,33 and p,3 have unknown regular-
ity; thus an additional equation is required for determining the regularity of
(u,33 , p,3 ). This additional equation comes from the divergence-free condi-
tion: differentiating divu = 0 in x3 , we find that

u1 ,13 +u2 ,23 +u3 ,33 = 0. (9.45)

Equations (9.44) and (9.45) together then form a system of four equations
with four unknowns (or to be more precise, four functions with unknown
regularity), and one can solve for (u,33 , p,3 ) and obtain regularity informa-
tion.
Solving for u1 ,33 and u2 ,33 from (9.44a,b), estimates (9.33), (9.36), (9.39)
and (9.43) imply that u1 ,33 and u2 ,33 both belong to L2 (Ω), and u1 , u2 satisfy

ku1 kH 2 (Ω) + ku2 kH 2 (Ω) ≤ Ckf kL2 (Ω) . (9.46)

Moreover, (9.45) implies that u3 ,33 = −u1 ,13 −u2 ,23 ; thus u3 satisfies

ku3 kH 2 (Ω) ≤ Ckf kL2 (Ω) . (9.47)

Finally, (9.44c) implies that p,3 = f 3 + u3 ,11 +u3 ,22 +u3 ,33 ; thus
h i
kp,3 kL2 (Ω) ≤ C kf kL2 (Ω) + kukH 2 (Ω) ≤ Ckf kL2 (Ω) ,

which, together with (9.36) and (9.43), further implies that

kpkH 1 (Ω) ≤ Ckf kL2 (Ω) .

Now suppose that Ω is a general bounded C 2 -domain. The change of


variable x 7→ θm (x) allows us to rewrite −∆u + Dp = f as

−(Ak` Aj` umi ,j ),k +Aki pm ,k = f i ◦ θm in B+ (0, rm );

thus rearranging terms suggests that

−An` An` umi ,nn+Ani pm ,n = f i ◦ θm −Aαi q,α +Ak` Aj`,k umi ,j + Ak` Aj` umi ,jk .
X

(j,k)6=(n,n)

Let F i be the right-hand side. Then by (9.39) and (9.43),

kξm F kL2 (B+ (0,rm )) ≤ Ckf kL2 (Ω) . (9.48)


§9.3 The Stokes equations 381

As in the proof of the elliptic regularity in Section 5.9, we observe that


An` An` ≥ Ann = eT
n Aen ≥ λ > 0. Therefore, for i = 1, · · · , n,

Ani m Fi
−umi ,nn + p ,n = in B+ (0, rm ). (9.49)
An` An` An` An`

On the other hand, by Aki umi ,k = 0 in B+ (0, rm ),

Akα umα ,k +Aβn umn ,β +Ann umn ,n = 0 in B+ (0, rm );

thus rearranging terms implies that


h i
Ann umn ,n +Anα umα ,n = − Aβα umα ,β +Aγn umn ,γ .

Differentiating the identity above in xn , we find that

Ann umn ,nn +Anα umα ,nn


h i (9.50)
= − Aβα umα ,β +Aγn umn ,γ ,n −Ann,n umn ,n −Anα,n umα ,n .

Let G be the right-hand side of (9.50). Then by (9.39),

kξm GkL2 (B+ (0,rm )) ≤ Ckf kL2 (Ω) . (9.51)

Equations (9.49) and (9.50) form an (n + 1) × (n + 1) system

(An` An` )−1 An1 um1 ,nn (An` An` )−1 F 1


    
−1 0 ··· 0
.. .. .. ..
(An` An` )−1 An2
    
 0 . 0 .  .   . 
.. .. .. ..
    
.. = .
.
 
 . 0 0 .
 .   . 
n n −1 n mn (A` A` )−1 F n
n n
    
 0 ··· 0 −1 (A` A` ) An   u ,nn   
An1 ··· ··· Ann 0 p m ,n G

We note that the determinant of the (n+1)×(n+1) matrix on the left-hand


side is (−1)n ; thus (um1 ,nn , · · · , umn ,nn , pm ,n ) can be uniquely determined.
Moreover, since Ak` ∈ W 1,∞ (B+ (0, rm )), by (9.48) and (9.51) we obtain that
n
X
kξm D2 um kL2 (B+ (0,rm )) + kξm Dpm kL2 (B+ (0,rm )) ≤ Ckf kL2 (Ω) . (9.52)
k=1

Combining (9.33), (9.36) and (9.52) together, we finally conclude that (9.29)
for the case m = 1. 
382 CHAPTER 9. Fluid dynamics

Remark 9.20. As suggested by (9.31), the convolution-by-layers opera-


tor Λ has to act on a divergence-free vector in order to obtain the esti-
mate (9.32), given that p ∈ L2 (Ω). Therefore, in the proof of boundary
2 (Λ umj ),
  
estimates of Theorem 9.19, instead of using v = Λ ξm  α ,α ◦
−1 as a test function (the test function from Section 5.9), we use v i =
θm
 i 2 Λ (A` umj ),
  −1
θm,` Λ ξm  j α ,α ◦ θm as a test function in (9.27), because the

vector w` ≡ A`j umj is divergence-free; that is,

divw = (A`j umj ),` = A`j umj ,` = (divu) ◦ θm = 0.

This is essential for estimating (p, divv)L2 (Ω) .

Now suppose that (9.22b) is replaced by divu = g for some g ∈ H 1 (Ω),


and the boundary condition (9.22c) is replaced by u = h on ∂Ω × (0, T ) for
some h ∈ H 1.5 (∂Ω) satisfying
Z Z
g dx = h · NdS .
Ω ∂Ω

By Corollary 8.5, there exists a function w ∈ H 2 (Ω) satisfying divw = g,


 
w = h on ∂Ω, and kwkH 2 (Ω) ≤ C kgkH 1 (Ω) +khkH 1.5 (∂Ω) . Letting v = u−w,
we find that v satisfies (9.22) except that the external force is replaced by
f − ∆w, and this process leads to the estimate
h i
kvkH 2 (Ω) ≤ Ckf − ∆wkL2 (Ω) ≤ C kf kL2 (Ω) + kgkH 1 (Ω) + khkH 1.5 (∂Ω) .

We have the following useful generalization:

Theorem 9.21. Suppose that ∂Ω is of class C m+1 for some m ∈ N. Then


for any f ∈ H m−1 (Ω), g ∈ H m (Ω) and h ∈ H m+0.5 (∂Ω) satisfying
Z Z
g dx = h · NdS,
Ω ∂Ω

there exists a unique solution (u, p) ∈ W m+1 × H m (Ω)/R to the steady




state Stokes problem

−∆u + Dp = f in Ω × (0, T ) ,
divu = g in Ω × (0, T ) ,
u=h on ∂Ω × (0, T ) .
§9.3 The Stokes equations 383

Moreover, (u, p) satisfies


h i
kukH m+1 (Ω) + kpkH m (Ω) ≤ C kf kH m−1 (Ω) + kgkH m (Ω) + khkH m+0.5 (∂Ω) .

Corollary 9.22. Let ∂Ω be of class C m for some m ∈ N. Then

H m (Ω; Rn ) ∩ H01 (Ω; Rn ) = W m ⊕ (H m (Ω; Rn ) ∩ G), (9.54)



where G = u ∈ H01 (Ω; Rn ) − ∆u = Dp for some p ∈ L2 (Ω)/R .


Proof. Let E and F be the left and right-hand side space of (9.54), respec-
tively. We prove that F is closed and dense in E to conclude the corollary.
Let un ∈ F and un → u in H m (Ω; Rn ) ∩ H01 (Ω). By the definition of F ,
un = vn + wn with vn ∈ W m and −∆wn = Dqn for some qn ∈ L2 (Ω)/R.
Let fn = −∆un = −∆vn + Dqn . Since fn ∈ H m−2 (Ω), and (vn , qn ) is
the solution to the homogeneous Stokes problem with forcing fn , (vn , qn ) ∈
W m × (H m−1 (Ω)/R) satisfying

kvn kH m (Ω) + kqn kH m−1 (Ω) ≤ Ckfn kH m−2 (Ω) .

Since un → u in H m (Ω; Rn ) ∩ H01 (Ω), fn = −∆un is uniformly bounded


in H m−2 (Ω). Therefore, vn is uniformly bounded in H m (Ω; Rn ) and qn is
uniformly bounded in H m−1 (Ω)/R; thus there exists a subsequence nj of n
such that vnj * v in H m (Ω; Rn ) and qnj * q in H m−1 (Ω)/R. On the other
hand, fn → −∆u in H m−2 (Ω; Rn ), and −∆un = −∆vn + Dqn implies that
−∆u = −∆v + Dq. This proves that F is closed.
To prove the density, let ` ∈ E 0 and `ϕ = 0 for all ϕ ∈ F , and we
show that ` ≡ 0. Since E is dense in H01 (Ω; Rn ), ` has a unique extension
`˜ ∈ H −1 (Ω; Rn ) and `˜ vanishes on W 1 ⊕ G. In particular, `˜ vanishes on W 1 ;
thus Lemma 9.17 implies that

˜ = (p, divϕ)L2 (Ω)


`ϕ ∀ ϕ ∈ H01 (Ω; Rn )

for some p ∈ L2 (Ω)/R. Let z ∈ H01 (Ω; Rn ) be the weak solution of −∆z =
Dp. Then z ∈ G, and by `˜ vanishes on G, we find that for all w ∈ G,

˜ = 0.
hz, −∆wiH01 (Ω) = h−∆z, wiH01 (Ω) = (p, divw)L2 (Ω) = `w

In other words, (divz, q)L2 (Ω) = hz, DqiH01 (Ω) = 0 for all q ∈ L2 (Ω)/R; thus
divz = 0. Therefore, z ∈ W 1 ∩ G = {0} which implies ` ≡ 0. 
384 CHAPTER 9. Fluid dynamics

An orthonormal basis of W 1

Let P denote the Leray projector. The Stokes operator A is defined by


A = −P∆. For f ∈ L2 (Ω; Rn ), we use the notation u = A−1 f to indicate
that u solves

−∆u + Dp = f in Ω,
divu = 0 in Ω,
u=0 on ∂Ω .

By the regularity theory of the Stokes problem, A−1 : W 1 → W 3 is bounded;


thus the Rellich theorem implies that A−1 : W 1 → W 1 is compact. More-
over, suppose that f, g ∈ W 1 , and u = A−1 f , v = A−1 g. Then for some
p ∈ H 2 (Ω),

(A−1 f, g)W 1 = (Du, Dg)L2 (Ω) = (−∆u, g)L2 (Ω)


(9.55)
= (−∆u + Dp, g)L2 (Ω) = (f, g)L2 (Ω) = (f, A−1 g)W 1 .

Therefore, A−1 : W 1 → W 1 is symmetric. By Theorem B.38, there exists


a sequence {ẽj }∞ 1 −1
j=1 ⊆ W , a collection of eigenfunctions of A , such that
{ẽj }∞ 1
j=1 forms an orthonomal basis of W . By the regularity theory for the
Stokes problem, ẽj ∈ W k for all k ∈ N.
Suppose that µj is the eigenvalue associated to ẽj . By (9.55),

µj = µj (ẽj , ẽj )W 1 = (A−1 ẽj , ẽj )W 1 = (ẽj , ẽj )L2 (Ω) > 0;

1/2
thus A−1 is a positive operator on W 1 . Let λj = µ−1
j , and define ej = λj ẽj .
Then {ej }∞ 1 2
j=1 is an orthogonal basis in W which is orthonormal in L (Ω).
Moreover,
Aej = λj ej , λj → ∞ as j → ∞,

or there exists pj ∈ H k (Ω) (for all k) such that

−∆ej + Dpj = λj ej λj → ∞ as j → ∞.

In the following discussion, whenever a basis of W 1 is used, it is this orthog-


onal basis we just construct.
9.4. INCOMPRESSIBLE EULER WITH FREE-BOUNDARY 385

Remark 9.23. Similar to the argument above, A−1 : L2 (Ω; Rn ) → L2 (Ω; Rn )


is compact. Moreover, if f, g ∈ L2 (Ω; Rn ), and u = A−1 f , v = A−1 g, then
for some p, q ∈ H 1 (Ω),

(A−1 f, g)L2 (Ω) = (u, −∆v + Dq)L2 (Ω) = (u, −∆v)L2 (Ω)
= (−∆u, v)L2 (Ω) = (−∆u + Dp, v)L2 (Ω) = (f, A−1 g)L2 (Ω)

which implies that A−1 : L2 (Ω; Rn ) → L2 (Ω; Rn ) is symmetric. Therefore, a


collection of eigenfunctions of A−1 in L2 (Ω; Rn ) constitutes an orthonormal
basis {ēj }∞ 2 n 1
j=1 of L (Ω; R ). However, unlike the case of a basis of W , the
associated eigenvalue of ēj could be zero; that is, there exists ēj 6= 0 such
that A−1 ēj = 0. In this case, there exists pj ∈ H 1 (Ω) such that

Dpj = ēj .

Those eigenfunctions associated with non-zero eigenvalues are also eigen-


functions in W 1 ; thus we establish the Hodge decomposition

L2 (Ω; Rn ) = W 1 ⊕L2 (Ω) Dp p ∈ H 1 (Ω) ,




where the closure of W 1 is taken under the L2 -norm.

9.4 Incompressible Euler with free-boundary


9.4.1 The Eulerian description

For 0 ≤ t ≤ T , the evolution of a three-dimensional incompressible fluid with


a moving free-surface is modeled by the incompressible Euler equations:

ut + u · Du + Dp = 0 in Ω(t) , (9.56a)
divu = 0 in Ω(t) , (9.56b)
p=0 on Γ(t) , (9.56c)
V(Γ(t)) = u · n (9.56d)
u = u0 on Ω(0) , (9.56e)
Ω(0) = Ω . (9.56f)

The open subset Ω(t) ⊂ R3 denotes the changing volume occupied by the
fluid, Γ(t) := ∂Ω(t) denotes the moving free-surface, V(Γ(t)) denotes normal
386 CHAPTER 9. Fluid dynamics

velocity of Γ(t), and n(t) denotes the exterior unit normal vector to the
free-surface Γ(t). The vector-field u = (u1 , u2 , u3 ) denotes the Eulerian
velocity field, and p denotes the pressure function. We use the notation
D = (∂1 , ∂2 , ∂3 ) to denote the gradient operator. We have normalized the
equations to have all physical constants equal to 1.
This is a free-boundary partial differential equation to determine the ve-
locity and pressure in the fluid, as well as the location and smoothness of the
a priori unknown free-surface. A recent explosion of interest in the analysis
of the free-boundary incompressible Euler equations, particularly in irrota-
tional form, has produced a number of different methodologies for obtaining
a priori estimates, and the accompanying existence theories have mostly re-
lied on the Nash-Moser iteration to deal with derivative loss in linearized
equations when arbitrary domains are considered, or complex analysis tools
for the irrotational problem with infinite depth.
To avoid the use of local coordinate charts necessary for arbitrary ge-
ometries, for simplicity, we will assume that the initial domain Ω at time
t = 0 is given by
Ω = T2 × (0, 1) , (9.57)

where T2 denotes the 2-torus, which can be thought of as the unit square
(0, 1)2 with periodic boundary conditions. This permits the use of one global
Cartesian coordinate system. We only allow the top boundary

Γ = {x3 = 1}

to move, while the bottom boundary is fixed with boundary condition

u3 = 0 on {x3 = 0} × [0, T ] .

Later, we will explain how to treat the case that the initial domain is an
arbitrary bounded, open subset of R3 with H 4 -class boundary.

Einstein’s summation convention

Repeated Latin indices i, j, k,, etc., are summed from 1 to 3; for example,
2
F,ii := i=1,3 ∂x∂i ∂xi .
P
9.4. INCOMPRESSIBLE EULER WITH FREE-BOUNDARY 387

The Lagrangian description

We transform the system (9.56) into Lagrangian variables. We let η(x, t)


denote the “position” of the fluid particle x at time t. Thus,

∂t η = u ◦ η for t > 0 and η(x, 0) = x

where ◦ denotes composition so that [u ◦ η](x, t) := u(η(x, t), t) . We set

v = u ◦ η (Lagrangian velocity),
q = p ◦ η (Lagrangian pressure),
A = [Dη]−1 (inverse of the deformation tensor),
J = det[Dη] (Jacobian determinant of the deformation tensor),
a = J A (cofactor of the deformation tensor).

Since div u = 0, we have that det Dη = 1, and hence the cofactor matrix of
Dη is equal to [Dη]−1 . Using Einstein’s summation convention, and using
∂F
the notation F,k to denote ∂xk , the kth-partial derivative of F for k = 1, 2, 3,
the Lagrangian version of equations (9.56) is given on the fixed reference
domain Ω by

vti + Aki q,k = 0 in Ω × (0, T ] , (9.58a)


divη v = 0 in Ω × (0, T ] , (9.58b)
q=0 on Γ × (0, T ] , (9.58c)
v3 = 0 on {x3 = 0} × (0, T ] , (9.58d)
(η, v) = (e, u0 ) in Ω × {t = 0} , (9.58e)

where e(x) = x denotes the identity map on Ω. Notice that the free suface
Γ(t) is given by
Γ(t) = η(t)(Γ) .

The Lagrangian divergence divη v = Aji v i ,j . Equation (9.58a) can be written


vector form as
vt + AT Dq = 0 in Ω × (0, T ] , (9.58a’)

where AT denotes the transpose of A. Solutions to (9.58) which are suf-


ficiently smooth to ensure that η(t) are diffeomorphisms, give solutions to
(9.56) via the change of variables indicated above.
388 CHAPTER 9. Fluid dynamics

9.4.2 The Lagrangian vorticity equation

We make use of the permutation symbol



 1, even permutation of {1, 2, 3},
εijk = −1, odd permutation of {1, 2, 3},
0, otherwise ,

and the basic identity regarding the ith component of the curl of a vector
field u:
(curl u)i = εijk uk ,j .

Defining curlη v = curl u ◦ η, the chain rule shows, by taking the curl of the
Euler equations (9.58a), that

(curlη vt )i = εijk asj vtj ,s = 0. (9.59)

9.4.3 Notation
Differentiation and norms

For integers k ≥ 0 and a smooth, open domain Ω of R3 , we define the Sobolev


space H k (Ω) (H k (Ω; R3 )) to be the completion of C ∞ (Ω) (C ∞ (Ω; R3 )) in
the norm
XZ
kuk2k := |Da u|2 dx,
|a|≤k Ω

where Da denotes all partial derivatives of order a. For real numbers s ≥ 0,


the Sobolev spaces H s (Ω) and the norms k · ks are defined by interpolation.
We will write H s (Ω) instead of H s (Ω; R3 ).
We define the horizontal derivative by ∂¯ = (∂1 , ∂2 ), and define the
Sobolev space H k (Γ) to be the completion of C ∞ (Γ) in the norm
XZ
|u|2 :=
k |∂¯a u|2 dS ,
|a|≤k Γ

where ∂¯a denotes all horizontal partial derivatives of order a, and dS =


dx1 dx2 denotes the ‘surface measure.’ For real s ≥ 0, the Hilbert space
H s (Γ) and the boundary norm |·|s is defined by interpolation. The negative-
order Sobolev spaces H −s (Γ) are defined via duality: for real s ≥ 0, H −s (Γ) :=
[H s (Γ)]0 .
9.4. INCOMPRESSIBLE EULER WITH FREE-BOUNDARY 389

The space of divergence-free vectors on Ω

In order to specify our initial velocity field, we introduce the following sub-
space of H s vector-fields on Ω for s ≥ 0:

Definition 9.24 (H s -class divergence free vectors).

s
Hdiv (Ω) = {u ∈ H s (Ω; R3 ) : u3 = 0 on {x3 = 0}, xh 7→ u(xh , x3 ) periodic
div u = 0} .

9.4.4 Properties of the cofactor matrix a, and a polynomial-


type inequality
Differentiating the inverse matrix A

Using that Dη A = Id, we have the following identities

¯ k = −As ∂η
∂A ¯ r , s Ak , (9.60)
i i r

DAki = −Asi Dη r ,s Akr , (9.61)


∂t Aki = −Asi v r ,s Akr . (9.62)

Relating the cofactor matrix and the unit normal n(t)

With N = (0, 0, 1) the outward unit normal to Γ, we have the identity

n(η) = aT N/|aT N | or ni (η) = a3i /|a3· | .

For α, β = 1, 2, we define the components of the induced metric on Γ(t) by

gαβ (x1 , x2 , t) = η,α (x1 , x2 , t) · η,β (x1 , x2 , t),

√ p
and let g denote det(gαβ ). The metric g is a 2 × 2 matrix defined on Γ.
It is an elementary computation to verify that


g = |η,1 ×η,2 | = |a3· | on Γ ,


and hence that a3i = gni (η).


A3i = J −1 gni (η) on Γ . (9.63)
390 CHAPTER 9. Fluid dynamics

9.4.5 Horizontal convolution-by-layers and commutation es-


timates
Horizontal convolution-by-layers
 
With xh = (x1 , x2 ), we define ρ ∈ C0∞ (R2 ) by ρ(xh ) = C exp 1
|xh |2 −1
if
|xh | < 1 and ρ(xh ) = 0 if |xh | ≥ 1; we then select the constant C so that
1 xh
R
R2 ρdxh = 1. We define ρκ (xh ) = κ2 ρ( κ ). It follows that for κ > 0,
0 ≤ ρκ ∈ C0∞ (R2 ) with spt(ρκ ) ⊂ B(0, κ). (Here, spt stands for support.)
We define the operation of horizontal convolution-by-layers as follows:
Z
Λκ f (xh , x3 ) = ρκ (xh − yh )f (yh , x3 )dyh for f ∈ L1 (R2 ) .
R2
By standard properties of convolution, there exists a constant C which
is independent of κ, such that for s ≥ 0,

kΛκ F ks ≤ CkF ks ∀ F ∈ H s (Ω) ,

and
|Λκ F |s ≤ C|F |s ∀ F ∈ H s (Γ) .
Furthermore,
¯ κ F k0 ≤ CkF k0
κk∂Λ ∀ F ∈ L2 (Ω) . (9.64)

Commutation estimates

¯ ∈ L2 (Γ), there is a generic


Lemma 9.25. For f ∈ W 1,∞ (Γ) and g, ∂g
constant C independent of κ such that
¯ − f Λκ ∂g|
|Λκ (f ∂g) ¯ 0 ≤ C|f |W 1,∞ (Γ) |g|0 ,

where W 1,∞ (Γ) denotes the Sobolev space of functions u ∈ L∞ (Γ) with weak
¯ ∈ L∞ (Γ).
derivative ∂u

Proof. We have that


1 xh − yh
Z
¯ − f Λκ ∂g](x)
[Λκ (f ∂g) ¯ = ρ( ¯ h )dyh
)[f (yh ) − f (xh )]∂g(y
B(xh ,κ) κ2 κ
1 ¯ x h − yh
Z
= ∂ρ( )[f (yh ) − f (xh )]g(yh )dyh
B(xh ,κ) κ3 κ
| {z }
I1
Z
− ¯ (yh )g(yh )dyh ,
ρκ (xh − yh )∂f
B(xh ,κ)
| {z }
I2
9.4. INCOMPRESSIBLE EULER WITH FREE-BOUNDARY 391

where we have used integration-by-parts in order to obtain the second equal-


ity. From Morrey’s inequality, for all yh ∈ B(xh , κ),

¯ kL∞ (T2 ) ≤ κ|f |W 1,∞ (Γ) ,


|f (xh ) − f (yh )| ≤ κk∂f

1 ¯ xh
so that with K(xh ) = κ2
|∂ρ( κ )|,

|I1 | ≤ |f |W 1,∞ (Γ) K ∗ |g| ,

so by Young’s inequality for convolution integrals,

|I1 |0 ≤ |f |W 1,∞ (Γ) kKkL1 (T2 ) |g|0 .

Similarly,
|I2 |0 ≤ |f |W 1,∞ (Γ) kρκ kL1 (T2 ) |g|0 .

The assertion is proved, given that

1 ¯ yh
Z Z
kKkL1 (T2 ) = |∂ρ( )|dyh = ¯
|∂ρ(z)|dz < ∞.
2
T2 κ κ T2

Lemma 9.26. For κ > 0, there exists C > 0 independent of κ, such that
1
for any g ∈ H 2 (Ω) and f ∈ H 3 (Ω), we have that

Λκ (f g) − f Λκ g 1 ≤ Cκkgk 1 kf k3 + Cκ 12 kgk0 kf k3 .

2 2

Proof. Let ∆ = Λκ (f g) − f Λκ g. Then, we have that


Z
∆(x) = ρκ (xh − yh )[f (yh , x3 ) − f (xh , x3 )] g(yh , x3 ) dyh .
B(xh ,κ)

Using the fact that H 2 (Ω) is embedded in L∞ (Ω), we have that


Z
|∆(x)| ≤ Cκkf k3 ρκ (xh − yh ) |g(yh , x3 )| dyh ,
B(xh ,κ)

showing that

k∆k0 ≤ Cκkf k3 kΛκ |g|k0


≤ Cκkf k3 kgk0 . (9.65)
392 CHAPTER 9. Fluid dynamics

Now, for j = 1, 2, 3,

∆,j = Λκ (f g,j ) − f Λκ g,j +Λκ (f,j g) − f,j Λκ g.

The difference between the two first terms of the right-hand side of this
identity can be treated in a similar fashion as (9.65), leading us to:

k∆,j k0 ≤ Cκkf k3 kgk1 + kΛκ (f,j g)k0 + kf,j Λκ gk0


≤ Cκkf k3 kgk1 + kf,j gk0 + kf,j kL∞ (Ω) kΛκ gk0
≤ Cκkf k3 kgk1 + 2kf,j kL∞ (Ω) kgk0
≤ Cκkf k3 kgk1 + Ckf k3 kgk0 . (9.66)

Consequently, we obtain by interpolation from (9.65) and (9.66):


1
k∆k 1 ≤ Cκkf k3 kgk 1 + Cκ 2 kf k3 kgk0 .
2 2

Following the estimate (9.66), we can similarly obtain estimates for


k∆,jk k0 and k∆,jkl k0 , j, k, l = 1, 2, 3, and thus estimate the H 2 (Ω)-norm as
well as the H 3 (Ω)-norm of ∆. Interpolation then yields the following:
3
Lemma 9.27. For κ > 0 and s = 2 or 25 , there exists C > 0 independent
of κ, such that for any g ∈ H s (Ω) and f ∈ H 3 (Ω), we have that

Λκ (f g) − f Λκ g ≤ Cκkgks kf k3 + Cκ 12 kgk 1 kf k3 .

s s− 2

9.4.6 An asymptotically consistent κ-approximation of the


Euler equations
The smoothed flow map ηκ

Definition 9.28 (The horizontally-smoothed flow map ηκ ). Suppose the


Lagrangian flow map η is in C([0, T ], H 3 (Ω)). For κ ∈ (0, κ0 ), we set on
[0, T ]
ηκ = Λ2κ η and vκ = ∂t ηκ .

By choosing T > 0 and κ0 sufficiently small, we can ensure that

Aκ (x, t) = [Dηκ (x, t)]−1

is well-defined on [0, T ]. We then define Jκ = det[Dηκ ], and aκ = Jκ Aκ .


9.4. INCOMPRESSIBLE EULER WITH FREE-BOUNDARY 393

Smoothing the initial data

In order to construct solutions to (9.58), we will introduce an approximation


scheme below, and it will be convenient to smooth the initial velocity field.
Let 0 ≤ %κ ∈ C0∞ (R3 ) denote the standard family of mollifiers with
spt(%κ ) ⊂ B(0, κ), and let EΩ denote the Sobolev extension operator map-
ping H s (Ω) to H s (R3 ) for s ≥ 0.
We set w0κ = %κ EΩ (u0 ), and define

uκ0 = w0κ − Drκ ,

where the scalar function rκ is the solution to the elliptic problem

∆rκ = div w0κ in Ω , (9.67a)


xh 7→ rκ (xh , x3 ) periodic x3 ∈ [0, 1] , (9.67b)
rκ = 0 on Γ , (9.67c)
∂rκ
= (w0κ )3 on {x3 = 0} . (9.67d)
∂x3
4 (Ω), it follows that uκ ∈ H s (Ω) for all s ≥ 4, and
Thus, given u0 ∈ Hdiv 0 div
that uκ0 → u0 in H 4 (Ω) as κ → 0.

The approximate κ-problem

For κ ∈ (0, κ0 ), we consider the following sequence of approximate problems:

vt + ATκ Dq = 0 in Ω × (0, Tκ ] , (9.68a)


divηκ v = 0 in Ω × (0, Tκ ] , (9.68b)
q=0 on Γ × (0, Tκ ] , (9.68c)
v3 = 0 on {x3 = 0} × (0, Tκ ] , (9.68d)
(v, η) = (uκ0 , eκ ) on Ω × {t = 0} , (9.68e)

where the solution η = η(κ) depends on the parameter κ, but for nota-
tional simplification, we do not explicitly write this dependence, and where
divηκ v = (Aκ )ji v i ,j , and eκ = Λ2κ e, where e(x) = x denotes the identity map
on Ω. We refer to the approximation (9.68) as the κ-problem. (Note that
our solution is periodic in the x1 and x2 directions.)
394 CHAPTER 9. Fluid dynamics

9.4.7 Construction of smooth solutions to κ-approximate Eu-


ler equations

Definition 9.29 (The manifold of volume-preserving embeddings). For s ≥


3, we let

s
Dvol = {η ∈ H s (Ω; R3 ) : η({x3 = 0}) ⊂ {x3 = 0}, xh 7→ η(xh , x3 ) periodic
det Dη = 1 , η −1 ∈ H s (η(Ω); Ω) } .

s is a infinite-dimensional C ∞ Hilbert manifold. We let


For s ≥ 3, Dvol
s denote the tangent bundle over D s , and T T D s the second tangent
T Dvol vol vol
s consist of pairs (η, v) ∈ D s ×H s (Ω)◦η,
bundle. Locally, elements of T Dvol vol div
where

s s
Hdiv (Ω) ◦ η = {u ◦ η : u ∈ Hdiv (Ω)} .

Theorem 9.30. Suppose that for s ≥ 4, uκ0 ∈ H s (Ω) with div u0 = 0.


Then for each κ ∈ (0, κ0 ), there exists a unique solution (η(κ), v(κ)) ∈
C ∞ ([0, Tκ ]; T Dvol
s ) to (9.68) with T > 0 depending on kuκ k an on κ > 0.
κ 0 s

Proof. With ηκ and vκ defined in Definition 9.28, we define

u = v ◦ ηκ−1 , uκ = vκ ◦ ηκ−1 , and p = q ◦ ηκ−1 .

It follows from the chain-rule that (9.68) can be written on the time depen-
dent domain ηκ (t, Ω) as

ut + uκ · Du + Dp = 0 in ηκ (t, Ω) , (9.69a)
div u = 0 in ηκ (t, Ω) , (9.69b)
p=0 on ηκ (t, Γ)] , (9.69c)
(u, η) = (uκ0 , e) on Ω × {t = 0} , (9.69d)

Taking the divergence of (9.69a), we see that p(t) satisfies

−∆p = ui ,j ujκ ,i in ηκ (t, Ω) , (9.70a)


p=0 on ηκ (t, Γ) , (9.70b)
∂p
=0 on {x3 = 0} . (9.70c)
∂N
9.4. INCOMPRESSIBLE EULER WITH FREE-BOUNDARY 395

The estimates for p(t) can be easily obtained by transforming (9.5), set on
the smoothed moving domain ηκ (t, Ω), to an elliptic equation on the fixed
domain Ω. It is important to note that this transformation should not
be made with the map ηκ (t), but rather with a family of diffeomorphisms
which inherits the smoothness of ηκ |Γ . To this end, consider the solution to
∆Φ(t) = 0 in Ω with Φ(t) = ηκ (t) on Γ. It follows that
1 1
kΦ(t)ks+1 ≤ C|ηκ |s+1/2 ≤ C(1 + )|η|s−1/2 ≤ C(1 + )kηks , (9.71)
κ κ
where the first inequality is the standard elliptic estimate for the Dirichlet
problem, the second follows from (9.64), and the third from the trace theo-
rem. For κ and Tκ taken sufficiently small kΦ(t)−eks can be made arbitrarily
small on [0, Tκ ], from which it follows that each such Φ(t) : Ω → ηκ (t, Ω) is
a diffeomorphism.
Next, define the matrix B = [DΦ]−1 and the pressure function Q = p◦Φ.
Using the chain-rule, (9.5) is transformed to

−Bij [Bik Q,k ],j = ui ,j ujκ ,i ◦ Φ



in Ω , (9.72a)
Q=0 on Γ , (9.72b)
Q,k Bik Bi3 = 0 on {x3 } = 0 , (9.72c)

for 0 ≤ t ≤ Tκ . Elliptic estimates (in conjunction with the Sobolev embed-


ding theorem) show that

kQ(t)ks+1 ≤ CP (kΦ(t)ks+1 )kui ,j ujκ ,i ◦Φks−1

where P is a polynomial function of its argument; hence, together with


(9.71), we have the estimate

kQ(t)ks+1 ≤ Cκ P (kη(t)ks ) · kvk2s ,

where the constant Cκ depends on κ > 0 and, in fact, blows-up as κ → 0.


Since p = Q ◦ Φ−1 , we see that

kDp(t) ◦ η(t)ks ≤ Cκ P (kη(t)ks , kvks ) . (9.73)

s by
We define the function F on T Dvol

F(η, v) = −Dp ◦ ηκ ,
396 CHAPTER 9. Fluid dynamics

and write (9.68a) as the coupled system of first-order ordinary differential


s :
equations written on T Dvol

∂t (η, v) = (v, F) ,
(η, v)|t=0 = (e, u0 ) .

s → T T D s continuously;
According to the estimate (9.73), (v, F) : T Dvol vol
moreover, we proved in establishing the inequality (9.15) that the map η 7→
s to T T D s so that
Dp ◦ η is Lipschiptz continuous from T Dvol vol

s s
(v, F) : T Dvol → T T Dvol is a Lipschitz map.

In fact, this map is C ∞ . Thus, the fundamental theorem of ordinary differ-


ential equations (or Picard iteration) shows that there exists a time Tκ > 0
depending on the initial data (and, of course, κ > 0) such that

(η, v) = (η(κ), v(κ)) ∈ C ∞ ([0, Tκ ]; T Dvol


s
)

is a unique solution of (9.68). Since, by definition Q = 0 on Γ, and

q = Q ◦ Φ−1 ◦ ηκ , (9.74)

it follows that q = 0 on Γ as well. 

9.4.8 Asymptotic estimates which are independent of the


smooth parameter κ

According to Theorem 9.30, we have unique solutions to (9.68),

(η(κ), v(κ)) ∈ C ∞ ([0, Tκ ]; T Dvol


s
),

and q(κ) given by (9.74). We will take s ≥ 6, and for notational convenience
we will denote (η(κ), v(κ)) by (η̃, ṽ), and write à for [Dη̃]−1 . We use the
notation η̃κ to denote Λ2κ η̃ and set Ãκ = [Dη̃κ ]−1 .

A continuous-in-time energy function appropriate for the asymp-


totic process κ → 0.

Definition 9.31. We set on [0, Tκ ]



Eκ (t) = 1 + kΛκ η̃(t)k24.5 + kṽ(t)k24 + kκṽ(t)k24.5 + kṽt (t)k23.5 + k κṽt (t)k24 .
(9.75)
9.4. INCOMPRESSIBLE EULER WITH FREE-BOUNDARY 397

The function Eκ (t) is the higher-order energy function which we will


prove remains bounded on a time-interval which is independent of κ. Given
6 , the E (t) is continuous on [0, T ].
(η̃, ṽ) ∈ T Dvol κ κ

Definition 9.32. We set E(t) = Eκ=0 (t) .

Definition 9.33. We set the constant M0 to be a polynomial function of


E(0) so that
M0 = P (E(0), k curl uκ0 k23.5 ) . (9.76)

Statement of the main result


4 , we obtain the initial pressure func-
Given an initial velocity field u0 ∈ Hdiv
tion p0 as the solution to the elliptic equation

−∆p0 = ui0 ,j uj0 ,i in Ω ,


p0 = 0 on Γ ,
p 0 ,3 = 0 on {x3 = 0} .

4 (Ω) with curl u ∈


Theorem 9.34 (Main Result). Given initial data u0 ∈ Hdiv 0

H 3.5 (Ω) such that

∂p0
− (x) ≥ λ > 0 for x ∈ Γ,
∂x3
there exists a solution to (9.58) verifying

sup E(t) ≤ P (E(0)) .


t∈[0,T ]

5 (Ω) and curl u ∈ H 4.5 (Ω), then the solution is unique.


If u0 ∈ Hdiv 0

Remark 9.35. The same theorem and proof hold in the case that Ω ⊂ R2 .
Below, we will treat the case of a general initial domain Ω.

Remark 9.36. The regularity for the existence theory is not optimal. In
fact, for our domain Ω, all that is necessary to establish existence and
3 (Ω);
uniqueness of solutions to (9.58) is an initial velocity field u0 ∈ Hdiv
nevertheless, the assumptions of Theorem 9.34 allow for the most transpar-
ent proof.
398 CHAPTER 9. Fluid dynamics

Conventions about constants


6 ,
As noted above, Theorem 9.30 provides us with solutions (η̃, ṽ) ∈ T Dvol
and hence supt∈[0,Tκ ] Eκ (t) is continuous.
We take Tκ > 0 sufficiently small so that, using the fundamental theorem
of calculus, for constants c1 , c2 and t ∈ [0, Tκ ],

−q̃,3 (t) ≥ λ/2 ,


c1 det g̃κ (0) ≤ det g̃κ (t) ≤ c2 det g̃κ (0) on Γ ,
c1 det J˜κ (0) ≤ det J˜κ (t) ≤ c2 det J˜κ (0) in Ω ,
kΛ̃κ η(t)k4 ≤ |Λκ e|4 + 1 , kq̃(t)k4 ≤ kq̃(0)k4 + 1 ,
kṽ(t)k3.5 ≤ ku0 k3.5 + 1 , kṽt (t)k3 ≤ kṽt (0)k3 + 1 .

The right-hand sides appearing in the last three inequalities shall be denoted
by a generic constant C in the estimates that we will perform.

Curl and divergence estimates for η̃, ṽ, and ṽt

Proposition 9.37. For all t ∈ (0, T ), with T ≤ Tκ ,



kcurl Λκ η̃(t)k23.5 + kcurl ṽ(t)k23 + kκcurl ṽ(t)k23.5 + k κcurl ṽt (t)k22.5
≤ M0 + C T P ( sup Eκ (t)) . (9.77)
t∈[0,T ]

Proof. By taking the curl of (9.68a), we have that

curlηκ ṽt = 0 .

From (9.59), it follows that (curlη̃κ ṽ)kt = B(Ãκ , Dṽ), where

s
B(Ãκ , Dṽ) = εkji (Ãκ )t j ṽ i ,s = εkij ṽ i ,s (Ãκ )sp ṽκp ,l (Ãκ )lj ;

hence,
Z t
curlη̃κ ṽ(t) = curl uκ0 + B(Ãκ (t0 ), Dṽ(t0 ))dt0 . (9.78)
0

Step 1. Estimate for curl Λκ η̃. Computing the gradient of (9.78) yields
Z t
curlη̃κ Dṽ(t) = D curl uκ0 − ε·ji D(Ãκ )sj ṽ i ,s + DB(Ãκ (t0 ), Dṽ(t0 ))dt0 .
0
9.4. INCOMPRESSIBLE EULER WITH FREE-BOUNDARY 399

Applying the fundamental theorem of calculus once again, shows that


Z t
s
curlη̃κ Dη̃(t) = tD curl uκ0 + ε·ji [(Ãκ )t j Dη̃ i ,s −D(Ãκ )sj ṽ i ,s ]dt0
0
Z tZ t0
+ DB(Ãκ (t00 ), Dṽ(t00 ))dt00 dt0 ,
0 0
and finally that
Z t
s
D curl η̃(t) = tD curl uκ0 − ε·ji (Ãκ )t j (t0 )dt0 Dη̃ i ,s (9.79)
0
Z t Z tZ t0
s
+ ε·ji [(Ãκ )t j Dη̃ i ,s −D(Ãκ )sj ṽ i ,s ]dt0 + DB(Ãκ (t00 ), Dṽ(t00 ))dt00 dt0 .
0 0 0

Using the fact that ∂t (Ãκ )sj = −(Ãκ )sl ṽκl ,p (Ãκ )pj and D(Ãκ )sj = −(Ãκ )sl Dη̃κl ,p (Ãκ )pj ,
we see that

DB(Ã, Dṽ) = −εkji [Dṽ i ,s (Ãκ )sl ṽκl ,p (Ãκ )pj + ṽ i ,s (Ãκ )sl Dṽκl ,p (Ãκ )pj
+ ṽ i ,s ṽκl ,p D((Ãκ )sl (Ãκ )pj )] .

The precise structure of the right-hand side is not very important; rather,
the derivative count is the focus, and as such we write

DB(Ã, D̃v) ∼ D2 ṽ Dṽκ Ãκ Ãκ + D2 ṽκ Dṽ Ãκ Ãκ + D2 η̃κ Dṽ Dṽκ Ãκ Ãκ .

Integrating by parts in time in the last term of the right-hand side of


(9.79), we see that
Z tZ t0 Z tZ t0 h i
DB(Ãκ , Dṽ) dt00 dt0 ∼ − D2 η̃ (Dṽκ Ãκ Ãκ )t + D2 η̃κ (Dṽ Ãκ Ãκ )t dt00 dt0
0 0 0 0
Z tZ t0
+ D2 η̃κ Dṽ Dṽκ Ãκ Ãκ dt00 dt0
0 0
Z th i
+ D2 η̃ Dṽκ Ãκ Ãκ + D2 η̃κ Dṽ Ãκ Ãκ dt0 .
0
Thus, we can write
Z t Z t
0
D curl η̃(t) ∼ tD curl uκ0 + D η̃2
Dṽκ Ãκ Ãκ dt + D2 η̃Dṽκ Ãκ Ãκ dt0
0 0
| {z } | {z }
I1 I2
Z t Z tZ t0
+ D2 η̃κ Dṽ Ãκ Ãκ dt0 + D2 η̃κ Dṽ Dṽκ Ãκ Ãκ dt00 dt0
|0 {z } |0 0
{z }
I3 I4
Z tZ t0 Z tZ t0
+ D2 η̃κ (Dṽ Ãκ Ãκ )t dt00 dt0 + D2 η̃ (Dṽκ Ãκ Ãκ )t dt00 dt0 .
|0 0
{z } |0 0
{z }
I5 I6
400 CHAPTER 9. Fluid dynamics

Our goal is to estimate kD curl Λκ ηk22.5 , which in turn requires us to


estimate kΛκ Ii k22.5 for i = 1, ..., 6. We begin with i = 1:
Z t
2 2
kΛκ I1 k2.5 ≤ kD Λκ η̃ Dṽκ Ãκ Ãκ dt0 k22.5
0
 Z t  Z t
+ kΛκ D2 η̃ Dṽκ Ãκ Ãκ dt0 − Dṽκ Ãκ Ãκ dt0 Λκ D2 η̃k22.5
0 0

It is easy to see that


Z t
2
kD Λκ η̃ Dṽκ Ãκ Ãκ dt0 k22.5 ≤ CT P ( sup Eκ (t)) ,
0 t∈[0,T ]

and by Lemma 9.27


 Z t  Z t
kΛκ D2 η̃ Dṽκ Ãκ Ãκ dt0 − Dṽκ Ãκ Ãκ dt0 Λκ D2 η̃k22.5
0 0
Z t Z t
0 2
2 2 2
≤ Cκ kD η̃k2.5 k 2 2
Dṽκ Ãκ Ãκ dt k3 + CκkD η̃k2 k Dṽκ Ãκ Ãκ dt0 k23
0 0
≤ CT P ( sup Eκ (t))
t∈[0,T ]

The same type of commutation estimate shows that

kΛκ I2 k22.5 ≤ CT P ( sup Eκ (t)) .


t∈[0,T ]

For i = 3, 4, 5, we use that kΛκ Ii k2.5 ≤ kIi k2.5 and as H 2.5 (Ω) is a multi-
plicative algebra, we see that for i = 3, 4, 5,

kΛκ Ii k22.5 ≤ CT P ( sup Eκ (t)) .


t∈[0,T ]

Finally, we consider the case that i = 6:


Z tZ t0
kΛκ I6 k22.5 ≤ k D2 Λκ η̃ (Dṽκ Ãκ Ãκ )t dt00 dt0 k22.5
0 0
Z tZ t0  h i   
+k Λκ D2 η̃ (Dṽκ Ãκ Ãκ )t − Dṽκ Ãκ Ãκ D2 Λκ η̃ dt00 dt0 k22.5
0 0 t
≤ CT P ( sup Eκ (t)) ,
t∈[0,T ]

where we have used Lemma 9.27 for the last inequality together with the
1
fact that kκ 2 ṽt k24 is contained in the energy function Eκ (t). Therefore, we
have proven that

kD curl Λκ ηk22.5 ≤ M0 + CT P ( sup Eκ (t)) ,


t∈[0,T ]
9.4. INCOMPRESSIBLE EULER WITH FREE-BOUNDARY 401

and hence with curlη̃κ ṽt = 0, that

k curl Λκ ηk23.5 ≤ M0 + CT P ( sup Eκ (t)) .


t∈[0,T ]

Step 2. Estimate for curl ṽ. From (9.78),


Z t Z t
curl v(t) = curl uκ0 + B(Ãκ (t0 ), Dṽ(t0 ))dt0 − ε·jk ṽ k ,r ∂˜t (Aκ )rj dt0 .
0 0

As H 3 (Ω) is a multiplicative algebra, it follows that on [0, T ],

k curl ṽ(t)k23 ≤ M0 + CT P ( sup Eκ (t)) .


t∈[0,T ]

Similarly,
kκ curl ṽ(t)k23.5 ≤ M0 + CT P ( sup Eκ (t)) .
t∈[0,T ]

Step 3. Estimate for curl ṽt . From (9.78),


Z t
curl ṽt (t) = εikj ∂t (Ãκ )rj (t0 )dt0 ṽti ,r ,
0

from which it follows that on [0, T ],



k κ curl ṽt (t)k23 ≤ CT P ( sup Eκ (t)) .
t∈[0,T ]

Proposition 9.38. For all t ∈ (0, T ), with T ≤ Tκ ,



kdiv Λκ η̃(t)k23.5 + kdiv ṽ(t)k23 + kκdiv ṽ(t)k23.5 + k κdiv ṽt (t)k23
≤ M0 + C T P ( sup Eκ (t)) . (9.80)
t∈[0,T ]

Proof. Since (Ãκ )ji ṽ i ,j = 0, we see that

(Ãκ )ji Dṽ i ,j = −D(Ãκ )ji ṽ i ,j . (9.81)

Step 1. Estimate for div Λκ η̃. It follows that

[(Ãκ )ji Dη̃ i ,j ]t = ∂t (Ãκ )ji Dη̃ i ,j −D(Ãκ )ji ṽ i ,j

so that
Z t 
[(Ãκ )ji Dη̃ i ,j ](t) = ∂t (Ãκ )ji Dη̃ i ,j −D(Ãκ )ji ṽ i ,j dt0 ,
0
402 CHAPTER 9. Fluid dynamics

and hence
Z t Z t Z t
D div η̃(t) = ∂t (Ãκ )ji Dη̃ i ,j dt0 − D(Ãκ )ji ṽ i ,j dt0 − ∂t (Ãκ )ji dt0 Dη̃ i ,j .
|0 {z } |0 {z } |0 {z }
I1 I2 I3

Thus,
3
X
kD div Λκ η̃(t)k22.5 ≤ kΛκ Ii (t)k22.5 .
i=1
Using Lemma 9.27 in the same fashion as was used for Propostion 9.37, we
see that
3
X
kΛκ Ii (t)k22.5 ≤ CT P ( sup Eκ (t)) ,
i=1 t∈[0,T ]

from which it follows that

k div Λκ η̃(t)k23.5 ≤ CT P ( sup Eκ (t)) .


t∈[0,T ]

Step 2. Estimate for div ṽ. From (Ãκ )ji ṽ i ,j = 0, we see that
Z t
div ṽ(t) = − ∂t (Ãκ )ji dt0 ṽ i ,j . (9.82)
0
Hence, it is clear that

k div ṽ(t)k23 + kκ div ṽ(t)k23.5 ≤ CT P ( sup Eκ (t)) .


t∈[0,T ]

Step 3. Estimate for div ṽt . From (9.82),


Z t
div ṽt (t) = −∂t (Ãκ )ji ṽ i ,j − ∂t (Ãκ )ji dt0 ṽti ,j .
0
It follows that

k div ṽt (t)k22.5 + k κ div ṽt (t)k23 ≤ CT P ( sup Eκ (t)) .
t∈[0,T ]

Pressure estimates

Letting (Ãκ )ji ∂x∂ j act on (9.68a), for t ∈ [0, Tκ ], the pressure function q(x, t)
satisfies the elliptic equation
h i
−(Ãκ )ji (Ãκ )ki q̃,k ,j = ṽ i ,j (Ãκ )jr ṽκr ,s (Ãκ )si in Ω ,
q̃ = 0 on Γ ,
q̃,k (Ãκ )ki (Ãκ )3i = ṽ · ∂t (Ãκ )3· = 0 on {x3 = 0} .
9.4. INCOMPRESSIBLE EULER WITH FREE-BOUNDARY 403

Using our conventions of Section 9.4.8 concerning the generic constant


C, we have the standard elliptic estimate on [0, Tκ ]

kq̃(t)k4.5 ≤ CkΛκ η̃(t)k4.5 . (9.83)

Similarly by time-differentiating the above elliptic equation for q, elliptic


estimates show that

kq̃t (t)k24 ≤ C kΛκ η̃(t)k24.5 + kṽ(t)k24 .




Technical lemma

Our energy estimates require the use of the following


1 1
Lemma 9.39. Let H 2 (Ω)0 denote the dual space of H 2 (Ω). There exists a
positive constant C such that
1
¯ k
k∂F 1 ≤ C kF k 1 ∀F ∈ H 2 (Ω) .
H 2 (Ω)0 2

Proof. Integrating by parts with respect to the horizontal derivative yields


for all G ∈ H 1 (Ω),
Z Z
¯ G dx = −
∂F ¯ dx ≤ CkF k0 kGk1 ,
F ∂G
Ω Ω

which shows that there exists C > 0 such that

¯ kH 1 (Ω)0 ≤ CkF k0 .
∀F ∈ L2 (Ω), k∂F (9.84)

Interpolating with the obvious inequality

¯ kL2 (Ω) ≤ CkF k1


∀F ∈ H 1 (Ω), k∂F

proves the lemma. 

Energy estimates

In this section, we take T ∈ (0, Tκ ).

Proposition 9.40. For t ∈ [0, Tκ ],

|Λκ η̃ 3 (t)|24 + |ṽ 3 (t)|23.5 ≤ M0 + C T P ( sup Eκ (t)) . (9.85)


t∈[0,T ]
404 CHAPTER 9. Fluid dynamics

Proof. Taking the L2 (Ω) inner-produce ∂¯4 of (9.68a) with ∂¯4 ṽ i yields
1 d ¯4
Z Z
0= k∂ ṽ(t)k0 + ∂ (Ãκ )i q̃,k ∂ ṽ dx + (Ãκ )ki ∂¯4 q̃,k ∂¯4 ṽ i dx +R ,
2 ¯4 k ¯4 i
2 dt } |Ω
| {z {z } |Ω {z }
I1 I2 I3
(9.86)
where R denotes integrals consisting of lower-order terms which can easily
be shown, via the Cauchy-Schwarz inequality, to satisfy
Z T
|R(t)|dt ≤ M0 + C T P ( sup E(t)) .
0 t∈[0,T ]

Using the identity (9.60), we see that


Z
I2 = − (Ãκ )kr ∂¯4 η̃κr ,s (Ãκ )si q̃,k ∂¯4 ṽ i dx + R
ZΩ Z
k ¯4 r
= − (Ãκ )r ∂ ηκ (Ãκ )i q̃,k ∂ ṽ dxh + (Ãκ )kr ∂¯4 ηκr (Ãκ )si q,k ∂¯4 ṽ i ,s dx +R .
3 ¯4 i

|Γ {z } |Ω {z }
I2a I2b

Since q̃ = 0 on Γ, so that q̃,k = q̃,3 , we see that


Z
−I2a = (−q̃,3 )∂¯4 η̃κr (Ãκ )3r ∂¯4 ṽ i (Ãκ )3i dxh ,
Γ

Recalling that η̃κ = Λκ Λκ η̃,


Z
−I2a = (−q̃,3 )∂¯4 Λκ η̃ r (Ãκ )3r ∂¯4 Λκ ṽ i (Ãκ )3i dxh
|Γ {z }
K1
Z h   i
+ ∂¯4 Λκ η̃ r Λκ (−q̃,3 ) (Ãκ )3i (Ãκ )3r ∂¯4 ṽ i − (−q̃,3 ) (Ãκ )3i (Ãκ )3r Λκ ∂¯4 ṽ i dxh ,
|Γ {z }
K2

According to Lemma 9.25,


 
K2 (t) ≤ |∂¯4 Λκ η̃ r |0 Λκ (−q̃,3 ) (Ãκ )3i (Ãκ )3r ∂¯4 ṽ i − (−q̃,3 ) (Ãκ )3i (Ãκ )3r Λκ ∂¯4 ṽ i

0
≤ C|∂¯4 Λκ η̃ r |0 |q̃,3 (Ãκ )3i (Ãκ )3r |W 1,∞ (Γ) |∂¯3 ṽ i |0 .

By the Sobolev embedding theorem,

|q̃,3 (Ãκ )3i (Ãκ )3r |W 1,∞ (Γ) ≤ C|q̃,3 |1.5 |(Ãκ )3i |1.5 |(Ãκ )3r |1.5 ,

so that Z T
K2 (t)dt ≤ C T P ( sup Eκ (t)) .
0 t∈[0,T ]
9.4. INCOMPRESSIBLE EULER WITH FREE-BOUNDARY 405

To study the integral K1 , we define ñκ to be the outward unit normal to


the moving surface η̃κ (t, Γ), given by ñκ = (ãκ )3i /|(ãκ )3· |. From (9.63),

(Ãκ )3· = J˜κ−1


p
g̃κ nκ on Γ ,

where
p
g̃κ = |η̃κ ,1 ×η̃κ ,2 | = |(ãκ )3· | on Γ .

It follows that
Z
K1 (t) = (−q̃,3 )∂¯4 Λκ η̃ · ñκ ∂¯4 Λκ ṽ · ñκ | det g̃κ |J˜κ−2 dxh
Γ
1d
Z
= (−q̃,3 )|∂¯4 Λκ η̃ · ñκ |2 | det g̃κ |J˜κ−2 dxh
2 dt Γ
| {z }
K1a
1 ¯4
Z
− ∂ Λκ η̃ i ∂¯4 Λκ η̃ j ∂t [(ñκ )i (ñκ )j | det g̃κ | J˜κ−2 ]dxh .
Γ 2
| {z }
K1b

By the assumption of Section 9.4.8,



˜−2
∂t [(ñκ )i (ñκ )j | det g̃κ | Jκ ]

≤C,
L∞ (Γ)

from which it follows that


Z T
K1b (t)dt ≤ C T P ( sup Eκ (t)) .
0 t∈[0,T ]

Using our assumed bounds for −q̃,3 (t), det g̃κ (t), J˜κ on [0, Tκ ], we see that
Z T
¯4 2
c̄|∂ Λκ η̃(t) · ñκ (t)|0 ≤ K1 (t)dt + M0 + C T P ( sup Eκ (t)) ,
0 t∈[0,T ]

for a constant c̄ which depends on λ, g̃κ (0), J˜κ (0).


Rt
Notice that N = (0, 0, 1) = ñκ (t)− 0 ∂t ñκ (t0 )dt0 , and by our assumptions
in Section 9.4.8, |∂t ñκ (t)|L∞ (Γ) ≤ C, so that with η̃ 3 = η̃ · N ,
Z T
3
c̄|Λκ η̃ (t)|24 ≤ K1 (t)dt + M0 + C T P ( sup Eκ (t)) ,
0 t∈[0,T ]

and hence
Z T
3
c̄|Λκ η̃ (t)|24 ≤− I2a (t)dt + M0 + C T P ( sup Eκ (t)) ,
0 t∈[0,T ]
406 CHAPTER 9. Fluid dynamics

RT RT
It remains to show that the integrals 0 I2b (t)dt and 0 I3 (t)dt are both
bounded by C T P (supt∈[0,T ] Eκ (t)). Using (9.68b),
Z
I2b (t) = − (Ãκ )kr ∂¯4 η̃κr q̃,k ṽ i ,s ∂¯4 (Ãκ )si dx + R

¯4
≤ Ck∂ η̃κ (t)k 1 k∂¯4 Ãκ (t)k 1 +R
2 H 2 (Ω)0

≤ Ck∂¯4 η̃κ (t)k 1 k∂¯3 Ãκ (t)k 1 +R


2 H 2 (Ω)
≤ C sup E(t) + R ,
t∈[0,T ]

where we have used Lemma 9.39 for the second inequality.


Finally,
Z Z
I3 (t) = − ∂¯4 q̃ ∂¯4 ṽ i ,k (Ãκ )ki dx = ∂¯4 q̃ ṽ i ,k ∂¯4 (Ãκ )ki dx + R
Ω Ω
¯3
≤ Ck∂ q̃(t)k 1 k∂¯4 Ãκ (t)k 1 +R
2 H 2 (Ω)0
≤ C sup E(t) + R ,
t∈[0,T ]

where we have used the pressure estimate (9.83) and Lemma 9.39 for the
last inequality.
Summing the estimates for I1 , I2 , I3 and integrating (9.86) from 0 to
T , we obtain the inequality,

|η̃ 3 |24 + k∂¯4 ṽ(t)k24 ≤ M0 + C T P ( sup Eκ (t)) .



sup
t∈[0,T ] t∈[0,T ]

According to Proposition 9.38,

sup k div ṽ(t)k23 ≤ M0 + C T P ( sup Eκ (t)) ,


t∈[0,T ] t∈[0,T ]

from which it follows that

sup k∂¯4 div ṽ(t)k2H 1 (Ω)0 ≤ M0 + C T P ( sup Eκ (t)) .


t∈[0,T ] t∈[0,T ]

Hence, the normal trace estimate (9.1) shows that

|η̃ 3 (t)|24 + |ṽ 3 (t)|23.5 ≤ M0 + C T P ( sup Eκ (t)) .



sup (9.87)
t∈[0,T ] t∈[0,T ]


9.4. INCOMPRESSIBLE EULER WITH FREE-BOUNDARY 407

Combining Proposition 9.40 with the curl estimates in Proposition 9.37


and the divergence estimates in Proposition 9.38 for η(t) and v(t) and using
(??) provides the following

Proposition 9.41. For t ∈ [0, Tκ ],

kΛκ η̃(t)k24.5 + kṽ(t)k24 ≤ M0 + C T P ( sup Eκ (t)) . (9.88)


t∈[0,T ]

Since ṽt = −ÃTκ Dq̃, we also obtain

Proposition 9.42. For t ∈ [0, Tκ ],

kṽt (t)k23.5 ≤ M0 + C T P ( sup Eκ (t)) . (9.89)


t∈[0,T ]

Proposition 9.43. For t ∈ [0, Tκ ] and α = 1, 2,



| κṽtα (t)|23.5 + |κṽ α (t)|24 ≤ M0 + C T P ( sup Eκ (t)) . (9.90)
t∈[0,T ]

Proof. Multiplying (9.68a) by η̃κj , we find the identity

ṽt · η̃κ ,k = −q̃,k for k = 1, 2, 3 .

It follows that
¯ κ = −∂¯q̃ = 0 on Γ ,
ṽt · ∂η

so that the tangential component (with respect to the moving boundary


η̃κ (t, Γ)) of ṽt vanishes on Γ. Hence,

∂¯3 ṽt · ∂¯η̃κ = −ṽt · ∂¯4 η̃κ − 3∂¯2 ṽt · ∂¯2 η̃κ − 3∂ṽ
¯ t · ∂¯3 η̃κ .

Thus,
|∂¯3 ṽt · ∂¯η̃κ |0 ≤ C|ṽt |2.5 |η̃κ |4 ≤ C|η̃κ |4 ,

so that using the fundamental theorem of calculus, together with the fact
that ηκ ,α (0) is proportional to Tα for α = 1, 2, we find that

|ṽtα |3 ≤ C|η̃κ |4 + C T |ṽt |3 , (9.91)

Similarly,
|∂¯4 ṽt · ∂¯η̃κ |0 ≤ C|ṽt |3 |η̃κ |5 ≤ C|η̃κ |5 ,
408 CHAPTER 9. Fluid dynamics

from which it follows that

|ṽtα |4 ≤ C|η̃|5 + C T |ṽt |4 . (9.92)

Interpolation between (9.91) and (9.92) yields

|ṽtα |3.5 ≤ C|η̃κ |4.5 + C T |ṽt |3.5 .

Interpolation between the inequalities |η̃κ |4 ≤ C|Λκ η̃|4 and |η̃κ |5 ≤ Cκ |Λκ η̃|4

shows that κ|η̃κ |4.5 ≤ C|Λκ η̃|4 . It thus follows from (9.88) and (9.89) that

κ|ṽtα (t)|3.5 ≤ M0 + C T P ( sup Eκ (t)) .
t∈[0,T ]

A similar argument shows that

κ|ṽ α (t)|4 ≤ M0 + C T P ( sup Eκ (t)) .


t∈[0,T ]

Combining the estimate (9.90) together with the curl estimates in Propo-

sition 9.37 and divergence estimates in Proposition 9.38 for κṽ(t) and κṽt (t)
proves the following

Proposition 9.44. For t ∈ [0, Tκ ] and α = 1, 2,



k κṽt (t)k24 + kκṽ(t)k24.5 ≤ M0 + C T P ( sup Eκ (t)) . (9.93)
t∈[0,T ]

9.4.9 Proof of the Main Theorem


9.4.10 Time of existence and bounds independent of κ and
existence of solutions to (9.58)

Summing the inequalities provided by the above Propositions, we find that

sup Eκ (t) ≤ M0 + C T P ( sup Eκ (t)) .


t∈[0,T ] t∈[0,T ]

A standard continuation argument then provides us with a time of ex-


istence T1 independent of κ and an estimate on (0, T1 ) independent of κ of
the type:
sup Eκ (t) ≤ 2M0 . (9.94)
t∈[0,T1 ]

In particular, our sequence of solutions (η̃, ṽ) satisfy the κ-independent


bound (9.94) on the κ-independent time-interval (0, T1 ).
9.4. INCOMPRESSIBLE EULER WITH FREE-BOUNDARY 409

9.4.11 The limit as κ → 0

By the κ-independent estimate (9.94), there exists a subsequence of {ṽt , Ãκ , Dq̃}
which converges uniformly to (vt , A, Dq) where A = [Dη]−1 , and η =
Rt
e + 0 vdt0 . Standard arguments show that (η, v) solve (9.58), and that

sup E(t) ≤ P (E(0)) .


t∈[0,T1 ]

Uniqueness

Suppose that on [0, T1 ], (η 1 , v 1 , q 1 ) and (η 2 , v 2 , q 2 ) are both solutions of


5 (Ω) and curl u ∈ H 4.5 (Ω), and with both
(9.58) with initial data u0 ∈ Hdiv 0

q1 and q2 satisfying the stability condition −∂qi /∂N > 0 on Γ for i = 1, 2.


Setting
Eη (t) = 1 + kη(t)k25.5 + kv(t)k25 ,

by the method of the previous section (with κ = 0), we infer that both Eη1 (t)
and Eη2 (t) are bounded by a constant M0 depending on the data u0 and Γ
on a time interval 0 ≤ t ≤ T1 for T1 small enough.
Let
w := v 1 − v 2 , r := q 1 − q 2 , and ξ := η 1 − η 2 .

Then (ξ, w, r) satisfies


Z t
ξ= w in Ω × (0, T ] , (9.95a)
0
∂t wi + (a1 )ki r,k = (a − a1 )ki q 2 ,k
2
in Ω × (0, T ] , (9.95b)
i
(a1 )ji wi ,j = (a2 − a1 )ji v 2 ,j in Ω × (0, T ] , (9.95c)
r=0 on Γ × (0, T ] , (9.95d)
(ξ, w) = (0, 0) on Ω × {t = 0} . (9.95e)

We set
E(t) = 1 + kξ(t)k24.5 + kw(t)k24 .

We will show that E(t) = 0, which shows that w = 0. We follow the


identical analysis as in the previous section, and estimate the new error
terms, arising from the difference of two solutions, using the additional space
regularity in our assumptions. With (9.95e), we find that supt∈[0,T ] E(t) ≤
C T P (supt∈[0,T ] E(t)).
410 CHAPTER 9. Fluid dynamics

Optimal regularity for initial data

We smoothed our initial data u0 in order to construct solutions to the κ-


approximation (9.68). Having obtained solutions which depend only on
E(0), a standard density argument shows that the initial data needs only
to satisfy E(0) < ∞. All assumptions on from Section 9.4.8 can now be
verified by the fundamental theorem of calculus, and taking T1 even small
if necessary.
Appendix A

A Brief Review on Analysis

A.1 Metric spaces

Let X be a set, and d : X × X → R be a real-valued function. (X, d) is


called a metric space if d satisfies

(1) d(x, y) ≥ 0 . d(x, y) = 0 if and only if x = y .

(2) d(x, y) = d(y, x) . (Symmetry)

(3) d(x, y) ≤ d(x, z) + d(z, y) . (Triangle inequality)

d is called a metric on X . d(x, y) is sometimes called the distance between


x and y .
A sequence xk ∈ X is said to converge to x ∈ X if lim d(xk , x) = 0 ,
k→∞
that is, given  > 0 , there exists N > 0 so that d(xk , x) <  for all k ≥ N .
A set U ⊆ X is said to be open in X if for all x ∈ U , there is an r > 0
so that the set

B(x, r) ≡ {y ∈ X | d(x, y) < r}

is a subset of U (that is, B(x, r) ⊆ U ). A set C ⊆ X is said to be closed in


X if X − C is open in X . A set K ⊆ X is said to be compact if every open
covering {Uα }α∈I of K possesses a finite subcovering {Uαk }K
k=1 .
The collection of all open sets in X is called the topology induced by
the metric d . In general, one can specify a collection T of subsets of X as
a topology on S if the following conditions hold:

411
412 APPENDIX A. A BRIEF REVIEW ON ANALYSIS

(1) ∅ ∈ T , X ∈ T ;
S
(2) If Uα ∈ T for α ∈ I , then Uα ∈ T ;
α∈I

K
T
(3) If Uk ∈ T for k = 1, 2, · · · , K , then Uk ∈ T .
k=1

The sets belonging to T are called open sets of X .

A.2 Linear spaces (Vector spaces)

A linear space consists of the following:

(1) a field of scalars F (usually R or C);

(2) a set X of objects, called vectors;

(3) a rule, called vector addition, which associates with each pair of vectors
x , y ∈ X , called the sum of x and y , in such a way that

(a) addition is commutative, that is, x + y = y + x ;

(b) addition is associative, that is, x + (y + z) = (x + y) + z ;

(c) there is a unique vector 0 in X , called the zero vector, such that
x + 0 = x for all x ∈ X ;

(d) for each vector x in X there is a unique y (which is usually denoted


by −x) in X such that x + y = 0 ;

(4) a rule, called scalar multiplication, which associates with each scalar c
in F and vector x in X a vector cx in X , called the product of c and
x , in such a way that

(a) 1x = x for every x ∈ X ;

(b) (c1 c2 )x = c1 (c2 x) ;

(c) c(x + y) = cx + cy ;

(d) (c1 + c2 )x = c1 x + c2 x .

The linear space X with field F is often called “a linear space X over F”.
A.3. NORMED SPACES 413

A.3 Normed spaces

A vector space X is said to be a normed space if to every x ∈ X there is


associated a non-negative reall number kxk , called the norm of x , in such
a way that

(1) kx + yk ≤ kxk + kyk for all x and y in X ;

(2) kcxk = |c|kxk if x ∈ X and c is a scalar;

(3) kxk > 0 if x 6= 0 .

A metric d on X can be induced by the norm k · k: d(x, y) = kx − yk .

A.4 Completeness

(1) Cauchy sequence: Let (X, d) be a metric space. A sequence {xk }∞


k=1 ∈
X is called a Cauchy sequence if for all ε > 0 , there exists a number
N > 0 so that d(xi , xj ) < ε for all i , j ≥ N .

(2) Complete metric spaces: A metric space (X, d) is complete if all


Cauchy sequence has a limit in X , that is, if {xk }∞
k=1 is a Cauchy
sequence, there is a point x ∈ X so that d(xk , x) → 0 as k → ∞ .

A.5 Completion of a metric space

Let (X, d) be a metric space but not complete. There exists a complete
˜ such that
metric space (X̃, d)

(1) there is an one-to-one map ι : X → X̃ ;

˜
(2) d(ι(x), ι(y)) = d(x, y) ; (isometry)

(3) ι(X) is dense in X̃ .

We can view X̃ as all possible limit points of (X, d) , and d˜ is “the” extension
of d .
If x ∈ X , we also use the notation x ∈ X̃ to denote that ι(x) ∈ X̃ . That
is, we identify x as an element in X̃ . However, we have to have in mind
414 APPENDIX A. A BRIEF REVIEW ON ANALYSIS

that conceptually this is not correct since every element in X̃ should be an


equivalent class of Cauchy sequence.
A complete normed linear space is called a Banach space.

A.6 Bounded linear maps between normed linear


spaces

Let X and Y be two linear spaces, and T : X → Y be linear, that is,

T (ax1 + bx2 ) = aT x1 + bT x2 ∀ a, b ∈ R(or C), x1 , x2 ∈ X .

If T : X → Y is ont-to-one and onto, then we can define the inverse map


T −1 : Y → X by T −1 y = x if and only if T x = y , so that T T −1 = T −1 T =
Id .

Definition A.1. Let (X, k · kX ) and (Y, k · kY ) be two normed linear spaces.
A linear map T : X → Y is bounded if there is a constant M ≥ 0 such that

kT xkY ≤ M kxkX ∀x ∈ X .

If no such constant exists, then we say that T is unbounded. If T : X → Y


is a bounded linear map, we define the operator norm of T by

kT k = inf{M | kT xkY ≤ M kxkX ∀ x ∈ X} .

We denote the set of all linear maps T : X → Y by L(X, Y ) , and the set
of all bounded linear maps T : X → Y by B(X, Y ) . When the domain and
range spaces are the same, we write L(X, X) = L(X) and B(X, X) = B(X) .
If in particular Y = R or C , then T is called a linear functional.

Note that
kT xkY
kT k = sup = sup kT xkY = sup kT xkY .
x6=0 kxkX kxkX ≤1 kxkX =1

Theorem A.2. A linear map is bounded if and only if it is continuous.

Definition A.3. Let X be a linear space. Two norms k · k1 and k · k2 on


X are equivalent if there are positive constants c and C such that

ckxk1 ≤ kxk2 ≤ Ckxk2 ∀x ∈ X .


A.7. CONVOLUTIONS AND INTEGRAL OPERATORS 415

Theorem A.4. Two norms on a linear space generate the same topology if
and only if they are equivalent.

Theorem A.5. All norms on Rn are equivalent.

Theorem A.6 (Open mapping theorem). Suppose that T : X → Y is


a one-to-one, onto bounded linear map between Banach space X and Y .
Then T −1 : Y → X is bounded.

A.7 Convolutions and integral operators

Suppose that f and g are two measurable functions on Rn . The convolution


of f and g is defined by
Z
(f ∗ g)(x) = f (y)g(x − y)dy x ∈ Rn .
Rn

Note that f ∗ g = g ∗ f , and the Tonelli theorem implies

kf ∗ gkL1 (Rn ) ≤ kf kL1 (Rn ) kgkL1 (Rn ) .

Theorem A.7 (Young’s inequality). Let p and q satisfy 1 ≤ p, q ≤ ∞ and


1 1 1 1 1
+ ≥ 1 , and let r be defined by = + − 1 . If f ∈ Lp (Rn ) and
p q r p q
g ∈ Lq (Rn ) , then f ∗ g ∈ Lr (Rn ) and

kf ∗ gkLr (Rn ) ≤ kf kLp (Rn ) kgkLq (Rn ) .

Proof. Use the identity


p p q q
|f (y)||g(x − y)| = |f (y)|1− r |f (y)| r |g(x−y)| r |g(x − y)|1− r
rp rq
and apply L r−p − Lr − L r−q type Hölder’s inequality. 

If u : Rn → R satisfies certain integrability conditions, then we define


the operator K acting on the function u as follows:
Z
(Ku)(x) = k(x, y)u(y)dy ,
Rn

where k : Rn × Rn → R is called the integral kernel. Given a function


f ∈ Lp (Rn ) , define an integral kernel k(x, y) = f (x − y) , then by Young’s
inequality, the corresponding integral operator K ∈ B(Lq (Rn ), Lr (Rn )) .
416 APPENDIX A. A BRIEF REVIEW ON ANALYSIS

A.8 The dual space and weak topology

A linear functional on a linear space X is a linear map from X to its scalar


fields. A continuous linear functional on a linear topological space (X ,T )
(with T denoting the topology which is called the strong topology) is a
continuous linear map from X to its scalar fields. Note that when talking
about the continuous linear functionals, we must have a strong topology to
begin with. The collection of continuous linear functionals on X is a linear
space, and is denoted by X 0 . X 0 is also called the dual space of X .

Example A.8 (Linear functional which is not continuous). Let X = C ∞ ([0, 1])
with sup-norm k · k∞ . Define the linear functional T : X → R by

T (f ) = f 0 (0) .
sin(n2 x)
Then T is not continuous. In fact, let fn (x) = n . Then fn → 0 in
(X, k · k∞ ) as n → ∞ , but T (fn ) 6→ 0 as n → ∞ .

The weak topology of X is the weakest/smallest topology on X so that


every element in X 0 is still continuous. The term “weak” is in the sense of
comparing to the “strong” topology. The weak topology of X is denoted by
Tw .
We can define a norm on X 0 as kf kX 0 = sup |f (x)| . Since X 0 with
kxkX ≤1
the topology induced by the dual space norm is a topological vector space,
we can talk about its dual space and weak topology. Every element x in X
can be considered as a continuous linear map which is defined by

x(f ) = f (x) .

In general, the dual space of X 0 is larger than X . For a reflexive space X ,


X 00 = X . The weak* topology of X 0 is the weakest/smallest topology on
X 0 so that every element in X is still continuous.

Theorem A.9 (Characteristics of weak and weak* convergence). Let (X, T )


be a topological vector space.

(1) A sequence {xk }∞


k=1 converges to x in the weak topology of X (this is
the same as saying that xk converges to x weakly in X), denoted by
xk * x in X , if and only if lim f (xk ) = f (x) for all f ∈ X 0 .
k→∞
A.9. BOUNDED LINEAR FUNCTIONAL ON LP (Ω) 417

(2) A sequence {fk }∞ 0


k=1 converges to f in the weak* topology of X , de-

noted by fk * f in X 0 , if and only if lim fk (x) = f (x) for all x ∈ X .
k→∞

Using the notation from the linear algebra, from now on the value of the
linear map A at x is denoted by Ax instead of A(x) .

Theorem A.10 (Banach-Alaoglu). If V is an open set containing 0 in a


topological vector space X and if

K = {Λ ∈ X 0 | |Λx| ≤ 1 ∀ x ∈ V } ,

then K is weak*-compact in X 0 .

A.9 Bounded linear functional on Lp (Ω)


p
Theorem A.11. Let p ∈ (1, ∞] , q = p−1 . For g ∈ Lq (Ω) , define Fg :
Lp (Ω) → R as
Z
Fg (f ) = f (x)g(x)dx .

Then Fg is a continuous linear functional on Lp (Ω) with operator norm


kFg kLp (Ω)0 = kgkLq (Ω) .

Theorem A.12 (Riesz Representation). Suppose that 1 ≤ p < ∞ and


p
φ ∈ Lp (Ω)0 . Then there exists g ∈ Lq (Ω) , q = p−1 such that
Z
φ(f ) = f (x)g(x)dx ∀ f ∈ Lp (Ω)

and kφkLp (Ω)0 = kgkLq (Ω) .

Corollary A.13. For p ∈ (1, ∞) , the space Lp (Ω) is reflexive, that is,
Lp (Ω)00 = Lp (Ω) .

Corollary A.14. If {fk }∞ p


k=1 is a bounded sequence in L (Ω) for some p ∈
(1, ∞) , then there exists a subsequence {fkj }∞ p
j=1 such that fkj * f in L (Ω) .

Corollary A.15. If {fk }∞ ∞


k=1 is a bounded sequence in L (Ω) , then there

exists a subsequence {fkj }∞ ∞
j=1 such that fkj * f in L (Ω) .
418 APPENDIX A. A BRIEF REVIEW ON ANALYSIS

Theorem A.16. If fk * f in Lp (Ω) as k → ∞ , then

kf kLp (Ω) ≤ lim inf kfk kLp (Ω) .


k→∞

Theorem A.17. If fk * f in Lp (Ω) as k → ∞ , then {fk }∞


k=1 is bounded
in Lp (Ω) .

Theorem A.18. Suppose that Ω ⊆ Rn is bounded, and

sup kfk kLp (Ω) ≤ M < ∞ and fk → f a.e.


k

If 1 < p < ∞ , then fk * f in Lp (Ω) .


Appendix B

Important Topics in
Functional Analysis

B.1 The Hahn-Banach theorem

Definition B.1. A vector space X is said to be a topological vector space if


there is a topology τ on X so such that

(a) every point of X is a closed set, and

(b) the vector space operations (addition of vectors and multiplication with
scalars) are continuous with respect to τ .

Definition B.2. The dual space of a topological vector space X is the vector
space X 0 whose elements are the continuous linear functionals on X .

Proposition B.3. A complex-linear functional on X is in X 0 if and only if


its real part is continuous, and that every continuous real-linear u : X → R
is the real part of a unique f ∈ X 0 .

Definition B.4. A map p from a real vector space V to R ∪ {±∞} is said


to be sub-linear over V if

p(λu) = λp(u) ∀ u ∈ V ,λ > 0,


p(u + v) ≤ p(u) + p(v) ∀ (u, v) ∈ V × V .

Theorem B.5. Let X be a real vector space, p a sub-linear function over


X , M a vector subspace of X . Suppose that T a linear functional over M

419
420APPENDIX B. IMPORTANT TOPICS IN FUNCTIONAL ANALYSIS

and T x ≤ p(x) on M . Then there exists a linear functional T̃ over X such


that

T̃ x = T x ∀x∈M,

and

−p(−x) ≤ T̃ x ≤ p(x) ∀ x∈X.

Corollary B.6. If X is a normed space and x0 ∈ X , there exists T ∈ X 0


such that

T x0 = kx0 kX and |T x| ≤ kxkX ∀ x∈X.

Theorem B.7. Let A and B are disjoint, non-empty, convex sets in a


topological vector space X .

(a) If A is open, then there exists T ∈ X 0 and γ ∈ R such that

Re T x < γ ≤ Re T y

for every x ∈ A and every y ∈ B .

(b) If A is compact, B is closed, and X is locally convex, then there exist


T ∈ X 0 , γ1 , γ2 ∈ R such that

Re T x < γ1 < γ2 < Re T y

for every x ∈ A and every y ∈ B .

Theorem B.8. Suppose M is a subspace of a locally convex space X , and


x0 ∈ X . If x0 is not in the closure of M , then there exists T ∈ X 0 such
that T x0 = 1 but T x = 0 for every x ∈ M .

Theorem B.9. If f is continuous linear functional on a subspace M of a


locally convex space X , then there exists T ∈ X 0 such that T = f on M .

Theorem B.10. Suppose B is a convex, balanced, closed set in a locally


convex space X , x0 ∈ X , but x0 6∈ B . Then there exists T ∈ X 0 such that
|T x| ≤ 1 for all x ∈ B , but T x0 > 1 .
B.2. THE OPEN MAPPING AND CLOSED GRAPH THEOREM 421

B.2 The open mapping and closed graph theorem

Theorem B.11 (The Baire Category Theorem). Let X be a complete metric


space.

(a) If {Un }∞
T
n=1 is a sequence of open dense subsets of X , then Un is
n=1
dense in X .

(b) X is not a countable union of nowhere dense sets.

Definition B.12 (Open mapping). Let X and Y be two topological vector


spaces. A mapping f : X → Y is said to be open if f (U ) is open in Y
whenever U is open in X .

Theorem B.13 (The Open Mapping Theorem). Suppose that X and Y be


Banach spaces, and T ∈ B(X, Y ) is surjective (i.e., onto). Then T is an
open mapping.

Corollary B.14 (The Bounded Inverse Theorem). Suppose that X and Y


be Banach spaces, and T ∈ B(X, Y ) is bijective (i.e., one-to-one and onto),
then the inverse map of T is bounded, or T −1 ∈ B(Y, X) . Equivalently,
there exist positive real numbers c and C such that

ckxkX ≤ kT xkY ≤ CkxkX ∀ x∈X.

Theorem B.15 (The Closed Graph Theorem). Suppose that X and Y are
Banach spaces, and T : X → Y is linear. If G = {(x, T x) | x ∈ X} is closed
in X × Y , then T ∈ B(X, Y ) .

B.3 Compact operators

Definition B.16 (Compact operators). Suppose X and Y are Banach


spaces and U is the open unit ball in X . A linear map T : X → Y is
said to be compact if the closure of T (U ) is compact in Y . It is clear that
T is then bounded. Thus T ∈ B(X, Y ) .

Definition B.17. An operator T ∈ B(X) is said to be invertible if there


exists S ∈ B(X) such that

ST = I = T S .
422APPENDIX B. IMPORTANT TOPICS IN FUNCTIONAL ANALYSIS

In this case, we write S = T −1 .

Definition B.18 (Spectrum and resolvent set). The specturm σ(T ) of an


operator T ∈ B(X) is the set of all scalars λ such that T −λI is not invertible,
and the resolvent set ρ(T ) is the complement of σ(T ) in the scalar field. Thus
λ ∈ σ(T ) if and only if at least one of the following two statements is true:

(i) The range of T − λI is not all of X .

(ii) T − λI is not one-to-one.

Definition B.19 (Classification of σ(T )). The spectrum of T ∈ B(X) is the


(disjoint) union of the following three sets:

(i) The point spectrum σp (T ) = {λ ∈ C | T − λI is not one-to-one} . If


λ ∈ σp (T ) , λ is also called an eigenvalue of T .

(ii) The continuous spectrum

σc (T ) = {λ ∈ C | T − λI is one-to-one, and has dense range}.

(iii) The residual spectrum

σr (T ) = {λ ∈ C | T − λI is one-to-one, and does not have dense range}.

Proposition B.20. The spectrum of a bounded operator T ∈ B(X) is


bounded.

Theorem B.21. Let X and Y be Banach spaces.

(a) If T ∈ B(X, Y ) and dim R(T ) < ∞ , then T is compact.

(b) If T ∈ B(X, Y ) , T is compact, and R(T ) is closed, then dim R(T ) <
∞.

(c) The compact operators form a closed subspace of B(X, Y ) in its norm-
topology.

(d) If T ∈ B(X) , T is compact, and λ 6= 0 , then dim N (T − λI) < ∞ .

(e) If dim X = ∞ , T ∈ B(X) , and T is compact, then 0 ∈ σ(T ) .


B.3. COMPACT OPERATORS 423

(f) If S ∈ B(X) , T ∈ B(X) , and T is compact, so are ST and T S .

Proof. (a) and (f) are trivial and left as exercises.

(b) If Y ≡ R(T ) is closed, then Y is complete, so that T is an open


mapping of X onto R(X) . Let U be the unit ball in X , then V ≡ T U
is open in Y . Since T is compact, V is pre-compact. Therefore, there
exist y1 , · · · , ym such that
m
[ 1
V ⊆ (yj + V ) . (1)
2
j=1

Let Z be the vector space spanned by y1 , · · · , ym . Then dim Z ≤ m ,


and Z is a closed subspace of Y . We also note that (1) implies V ⊆
Z + 21 V . Since Z = λZ for all λ 6= 0 ,

1 1 1
V ⊆Z + V ⊆Z +Z + V =Z + V .
2 4 4

We then see that



\
V ⊆ (Z + 2−n V ) = Z .
n=1

However, it would further implies that kV ⊆ Z for all k ∈ N , so


Z=Y .

(c) Let Σ be the collection of compact operators in B(X, Y ) , U be the


unit ball in X , and T ∈ Σ . For every r > 0 , there exists S ∈ Σ with
kS − T kB(X,Y ) < r . Since SU is totally bounded, there exists points
x1 , · · · , xn in U such that SU is covered by the balls B(Sxi , r) . Since
kSx − T xkY ≤ r for every x ∈ U , it follows that T U is covered by the
balls of B(T xi , 3r) . Thus T U is totally bounded as well, so T ∈ Σ .

(d) Let Y = N (T − λI) . The restriction of T to Y is a compact operator


whose range is Y . By (b), dim(Y ) < ∞ .

(e) T : X → X cannot be an onto map since if it is onto, then dim R(T ) =


∞ which contradicts to (b) . 
424APPENDIX B. IMPORTANT TOPICS IN FUNCTIONAL ANALYSIS

Definition B.22 (Adjoint operators). The adjoint operator T ∗ of an op-


erator T ∈ B(X, Y ) is the unique bounded operator belonging to B(Y 0 , X 0 )
satisfying

hT x, y ∗ iY = hx, T ∗ y ∗ iX .

Theorem B.23. Suppose X and Y are Banach spaces and T ∈ B(X, Y ) .


Then T is compact if and only if T ∗ is compact.

Proof. (⇒) Suppose T is compact. Let {yn∗ }∞


n=1 be a sequence in the unit
ball of Y 0 . Define

fn (y) = hy, yn∗ iY ∀y∈Y .

Since |fn (y1 )−fn (y2 )| ≤ ky1 −y2 kY , {fn }∞


n=1 is equi-continuous. Since T (U )
has compact closure in Y (as before, U is the unit ball of X), Arzela-Ascoli
theorem implies that {fn }∞ ∞
n=1 has a subsequence {fnj }j=1 that converges
uniformly on T (U ) . Since

kT ∗ yn∗ i − T ∗ yn∗ j kX 0 = sup |hT x, yn∗ i − yn∗ j iY | = sup |fni (T x) − fnj (T x)| ,
x∈U x∈U

the completeness of X 0 implies that {T ∗ yn∗ j }∞ ∗


j=1 converges. Hence T is
compact.
(⇐) can be proved in the same fashion. 

Definition B.24. Suppose M is a closed subspace of a topological vector


space X . If there exists a closed subspace N of X such that

X =M +N and M ∩ N = {0} ,

then M is said to be complemented in X . In this case, X is said to be the


direct sum of M and N , and the notation X = M ⊕ N is used.

Lemma B.25. Let M be a closed subspace of a Banach space X .

(a) If dim M < ∞ , then M is complemented in X .

(b) If dim(X/M ) < ∞ , then M is complemented in X .

The dimension of X/M is also called the codimension of M in X .


B.3. COMPACT OPERATORS 425

Proof. Note that the closedness of M is only used in (b), while in (a)
the closedness is implied by the finite dimensionality (so no assumption
is needed).

(a) Let {e1 , · · · , en } be a basis for M . Then every x ∈ M has a unique


representation

x = α1 (x)e1 + · · · + αn (x)en .

αi is a continuous linear functional which vanishes on the span of


{e1 , · · · , ei−1 , ei+1 , · · · , en } , and can be extended to a continuous lin-
ear functional that only take non-zero values in the 1-dimensional
space spanned by ei . Let N be the intersection of the null space
of these extensions. Then X = M ⊕ N .

(b) Let {e1 , · · · , en } be a basis of X/M (closedness of M is used to define


the quotient space), and π : X → X/M be the quotient map. Pick
xi ∈ X so that πxi = ei , and define N to be the span of {x1 , · · · , xn } .
Then X = M ⊕ N . 

Lemma B.26. Let M be a subspace of a normed space X . If M is not


dense in X , and if r > 1 , then there exists x ∈ X such that

kxkX < r but kx − ykX ≥ 1 ∀ y ∈ M.

Proof. There exists x1 ∈ X whose distance from M is 1, that is,

inf{kx1 − ykX |y ∈ M } = 1 .

Choose y1 ∈ M such that kx1 − y1 kX < r , and put x = x1 − y1 . 

Theorem B.27. If X is a Banach space, T ∈ B(X) , T is compact, and


λ 6= 0 , then T − λI has closed range.

Proof. By (d) of Theorem B.21, dim N (T − λI) < ∞ . By (a) of Lemma


B.25, X is the direct sum of N (T − λI) and a closed subspace M . Define
an operator S ∈ B(M, X) by

Sx = T x − λx .
426APPENDIX B. IMPORTANT TOPICS IN FUNCTIONAL ANALYSIS

Then S is one-to-one on M . Also, R(S) = R(T − λI) . Similar to the proof


of Lax-Milgram theorem 5.15, to show that R(S) is closed, it suffices to show
the existence of an r > 0 such that

rkxkX ≤ kSxkX ∀x∈M.

Suppose the contrary that for every r > 0 , there exists {xn } in M such that
kxn kX = 1 , Sxn → 0 , and (after passage to a subsequence) T xn → x0 for
some x0 ∈ X (by the compactness of T ). It follows that λxn → x0 . Thus
x0 ∈ M since M is a closed subspace, and

Sx0 = lim (λSxn ) = 0 .


n→∞

Since S is one-to-one, x0 = 0 . However, kxn kX = 1 for all n , and x0 =


lim λxn , hence kx0 kX = |λ| > 0 . 
n→∞

Corollary B.28. The continuous spectrum of a compact operator T ∈ B(X)


contains at most one point, namely 0.

Theorem B.29. Suppose X is a Banach space, T ∈ B(X) , T is compact,


r > 0 , and E is a set of eigenvalues λ of T such that |λ| > r . Then

(a) for each λ ∈ E , R(T − λI) 6= X , and

(b) E is a finite set.

Proof. We first show that either (a) or (b) is false then there exist closed
subspaces Mn of X and scalars λn ∈ E such that

M1 ( M2 ( M3 ( · · · , (1)
T (Mn ) ⊆ Mn for n ≥ 1 , (2)
(T − λn I)(Mn ) ⊆ Mn−1 for n ≥ 2 . (3)

Suppose (a) is false. Then R(T − λ0 I) = X for some λ0 ∈ E . Let


S = T − λ0 I , and define Mn = N (S n ) , i.e., the null space of S n . Since λ0
is an eigenvalue of T , there exists x1 ∈ M1 , x1 6= 0 . Since R(S) = X , there
is a sequence {xn }∞
n=1 in X such that Sxn+1 = xn , n = 1, 2, 3, · · · . Then

S n xn+1 = x1 6= 0 but S n+1 xn+1 = Sx1 = 0 .


B.3. COMPACT OPERATORS 427

Hence Mn is a proper closed subspace of Mn+1 . It follows that (1) to (3)


hold, with λn = λ0 .
Suppose (b) is false. Then E contains a sequence {λn } of distinct eigen-
values of T . Choose corresponding eigenvectors en , and let Mn be the
(finite-dimensional, hence closed) subspace of X spanned by {e1 , · · · , en } .
Since λn are distinct, {e1 , · · · , en } is a linearly independent set, so that
Mn−1 is a proper subspace of Mn . This gives (1). If x ∈ Mn , then

x = α1 e1 + · · · αn en ,

which shows that T x ∈ Mn and

(T − λn I)x = α1 (λ1 − λn )e1 + · · · + αn−1 (λn−1 − λn )en−1 ∈ Mn−1 .

Thus (2) and (3) hold.


Once we have closed subspace Mn satisfying (1) to (3), Lemma B.26
gives us vectors yn ∈ Mn , for n = 2, 3, 4, · · · , such that

kyn kX ≤ 2 and kyn − xkX ≥ 1 if x ∈ Mn−1 . (4)

If 2 ≤ m < n , define

z = T ym − (T − λn I)yn .

By (2) and (3), z ∈ Mn−1 . Hence (4) shows that

kT yn − T ym kX = kλn yn − zkX = |λn |kyn − λ−1


n zkX ≥ |λn | > r .

The sequence {T yn }∞
n=1 has therefore no convergent subsequences, although
{yn }∞
n=1 is bounded, contradicting to the compactness of T . 

Remark B.30. Let H denote a Hilbert space. T ∈ B(H) is said to be


normal if T T ∗ = T ∗ T . A much deeper result states that a normal operator
T ∈ B(H) is compact if and only if it satisfies the following two conditions:

(a) σ(T ) has no limit point except possibly 0 .

(b) If λ 6= 0 , then dim N (T − λI) < ∞ .


428APPENDIX B. IMPORTANT TOPICS IN FUNCTIONAL ANALYSIS

Theorem B.31 (The Fredholm Alternative). Suppose X is a Banach space,


T ∈ B(X) , and T is compact.

(a) If λ 6= 0 , then the four numbers

α = dim N (T − λI) α∗ = dim N (T ∗ − λI)


β = dim X/R(T − λI) β ∗ = dim X/R(T ∗ − λI)

are equal and finite.

(b) If λ 6= 0 and λ ∈ σ(T ) , then λ is an eigenvalue of T and of T ∗ .

(c) σ(T ) is compact, at most countable, and has at most one limit point,
namely, 0.

Proof. Suppose M0 is a closed subspace of a locally convex space Y , and k


is a positive integer such that k ≤ dim Y /M0 . Then there are vectors y1 ,
· · · , yk in Y such that the vector space Mi generated by M0 and y1 , · · · , yi
contains Mi−1 as a proper subspace. Each Mi is closed, and hence by The-
orem B.8, there are continuous linear functionals T1 , · · · , Tk on Y such
that Ti yi = 1 but Ti y = 0 for all y ∈ Mi−1 . These functionals are linearly
independent, so if Σ denotes the space of all continuous linear functionals
on Y that annihilate M0 , then

dim Y /M0 ≤ dim Σ .

Let S = T − λI . Apply this with Y = X , M0 = R(S) . Since R(S)


is closed, Σ = R(S)⊥ = N (S ∗ ) , so β ≤ α∗ . Next, take Y = X 0 with its
weak*-topology, and M0 = R(S ∗ ) . A result from functional analysis states
that R(S ∗ ) is weak*-closed. Since Σ consists of all weak*-continuous linear
functional on X 0 that annihilate R(S ∗ ) , Σ is isomorphic to ⊥ R(S ∗ ) = N (S) ,
hence β ∗ ≤ α .
Next we show that α ≤ β , and the same proof can be used to show that
α∗ ≤ β ∗ , so the proof of (a) (and hence (b) and (c)) is complete. Assume
the contrary that α > β . By (d) of Theorem B.21, α < ∞ . By Lemma
B.25, there exists closed subspaces E and F such that dim F = β and

X = N (S) ⊕ E = R(S) ⊕ F .
B.3. COMPACT OPERATORS 429

Every x ∈ X has unique representation x = x1 + x2 , with x1 ∈ N (S) ,


x2 ∈ E . Define π : X → N (S) by πx = x1 . It is easy to see (by the closed
graph theorem B.15) that π is continuous.
Since we assume that dim N (S) > dim F , there is a linear mapping φ of
N (S) onto F such that φx0 = 0 for some x0 6= 0 . Define

Φx = T x + φπx ∀ x∈X.

Then Φ ∈ B(X) . Since dim R(φπ) < ∞ , φπ is a compact operator, hence


so is Φ .
Observe that Φ − λI = S + φπ . If x ∈ E , then πx = 0 , so (Φ − λI)x =
Sx ; hence

(Φ − λI)(E) = R(S) .

If x ∈ N (S) , then πx = x , (Φ − λI)x = φx ; hence

(Φ − λI)(N (S)) = φ(N (S)) = F .

Therefore, R(Φ − λI) = R(S) + F = X . Moreover, λ is an eigenvalue of Φ


(with x0 as a corresponding eigenvector), and since Φ is compact, Theorem
B.29 states that R(Φ − λI) cannot be all of X . 

Corollary B.32. The residual spectrum of a compact operator T ∈ B(X)


contains at most one point, namely 0 . Moreover, σ(T ) = σp (T ) ∪ {0} .

Corollary B.33. Suppose that H is a Hilbert space, and T ∈ B(H) is


compact. Then R(T − λI) = N (T ∗ − λI)⊥ for all λ 6= 0 .

Remark B.34. If T ∈ B(X) is compact, then the injectivity of T − λI


implies the invertibility of T − λI if λ 6= 0 .

Remark B.35. A much deeper result states that the spectrum of a bounded
operator T ∈ B(X) is also compact.
430APPENDIX B. IMPORTANT TOPICS IN FUNCTIONAL ANALYSIS

B.3.1 Symmetric operators on Hilbert Spaces

Let H be a Hilbert space, and T ∈ B(H) . By Riesz representation theorem,


given a continuous linear functional y ∗ ∈ H0 , there exists y ∈ H such that

hh, y ∗ iH = (h, y)H ∀ h ∈ H.

In particular, let h = T x , and suppose the representation of T ∗ y ∗ is z , then

(x, z)H = hx, T ∗ y ∗ iH = hT x, y ∗ iH = (T x, y)H ∀ h ∈ H.

The element z ∈ H is denoted by T 0 y . In this case, T 0 is also called the


adjoint operator of T (and T 0 can be thought as the representation of T ∗ ).

Definition B.36 (Symmetry). The operator T ∈ B(H) is called symmetric


if T = T 0 .

Lemma B.37. Suppose that T ∈ B(H) be symmetric, and

m≡ inf (T u, u)H , M ≡ sup (T u, u)H .


kukH =1 kukH =1

Then σ(T ) ⊆ [m, M ] , and m, M ∈ σ(T ) .

Proof. Let λ > M . Then L : H → H0 defined by

hLu, ϕiH = (λu − T u, ϕ)H ∀ϕ∈H

is bounded and coercive: the boundedness is trivial, and the coercivity fol-
lows from that

hLu, uiH = (λu − T u, u)H ≥ (λ − M )kuk2H .

Therefore, by the Lax-Milgram theorem, L : H → H0 is one-to-one and onto,


so is λI − T since λI − T is the representation of L . Therefore, λ 6∈ σ(T ) .
Similarly, λ 6∈ σ(T ) if λ < m . So σ(T ) ⊆ [m, M ] .
Let [u, v] = (M u − T u, v)H . The proof of the Schwarz inequality (The-
orem 3.1) implies that

|[u, v]| ≤ |[u, u]|1/2 |[v, v]|1/2 .


B.3. COMPACT OPERATORS 431

Taking the supremum over all v such that kvkH = 1 , then


1/2
kM u − T ukH ≤ C(M u − T u, u)H ∀u∈H (1)

for some constant C .


Let {uk }∞
k=1 be such that kuk kH = 1 , and (T uk , uk )H → M . Then (1)
implies kM uk − T uk kH → 0 as k → ∞ . If M 6∈ σ(T ) , M I − T is invertible
and has a bounded inverse (by the bounded inverse theorem), so

uk = (M I − T )−1 (M uk − T uk ) → 0 in H

which contradicts to kuk kH = 1 for all k . Hence M ∈ σ(T ) . Similarly,


m ∈ σ(T ) . 

Theorem B.38. Let H be a separable Hilbert space, and suppose that T ∈


B(H) is compact and symmetric. Then there exists a countable orthonormal
basis of H consisting of eigenvectors of T .

Proof. Let {λk }∞


k=1 be the sequence of distinct eigenvalues of T , λk 6= 0 .
Set λ0 = 0 , and Hk = N (T − λk I) for k ≥ 0 . Then dim Hk < ∞ if k > 0 .
Moreover, if xi ∈ Hi and xj ∈ Hj , then

λi (xi , xj )H = (T xi , xj )H = (xi , T xj )H = λj (xi , xj )H ⇒ (xi , xj )H = 0 if i 6= j .

Therefore, the subspaces Hi and Hj are orthogonal.


Let H̃ be the smallest subspace of H consisting of all these Hi , i =
0, 1, · · · . Then
m
nX o
H̃ = ck uk m ∈ N ∪ {0} , uk ∈ Hk , ak ∈ R .

k=0

We note that T (H̃) ⊆ H̃ , and this further implies that T (H̃⊥ ) ⊆ H̃⊥ since

(T u, v)H = (u, T v)H = 0 ∀ u ∈ H̃⊥ , v ∈ H̃ .

The operator T̃ ≡ T |H̃⊥ , the restriction of T to H̃⊥ , is also compact and


symmetric. In addition, σ(T̃ ) = {0} , since any nonzero eigenvalue of T̃
would be an eigenvalue of T as well. According to the previous lemma,
(T̃ u, u)H = 0 for all u ∈ H̃ ⊥ . But then if u, v ∈ H̃ ⊥ ,

2(T̃ u, v)H = (T̃ (u + v), (u + v))H − (T̃ u, u)H − (T̃ v, v)H = 0


432APPENDIX B. IMPORTANT TOPICS IN FUNCTIONAL ANALYSIS

Hence T̃ = 0 on H̃⊥ . As a consequence, H̃⊥ ⊆ N (T ) ⊆ H̃ , so H̃⊥ = {0} .


Thus H̃ is dense in H .
An orthonormal basis of H then can be obtained by choosing an or-
thonormal basis for each subspace Hk , k = 0, 1, · · · . Note that the sepa-
rability of H implies that H0 has a countable orthonormal basis, and these
basis vectors are all eigenvectors corresponding to λ0 = 0 . 

You might also like